This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Optimum Location-Based Relay Selection in Wireless Networks

Hazer Inaltekin, , Saman Atapattu, ,
Jamie S. Evans, 
H. Inaltekin is with the School of Engineering, Macquarie University, North Ryde, NSW 2109, Australia. Email: [email protected]. Atapattu and J. S. Evans are with the Department of Electrical and Electronic Engineering, University of Melbourne, Parkville, VIC 3010, Australia. Email:{saman.atapattu, jse}@unimelb.edu.au.The material in this paper was presented in part at the IEEE International Conference on Communications, Shanghai, China, May 2019, and in part at the IEEE Global Communications Conference, Waikoloa, HI, USA, December 2019 [1, 2].This work has been submitted to the IEEE for possible publication. Copyright may be transferred without notice, after which this version may no longer be accessible.
Abstract

This paper studies the performance and key structural properties of the optimum location-based relay selection policy for wireless networks consisting of homogeneous Poisson distributed relays. The distribution of the channel quality indicator at the optimum relay location is obtained. A threshold-based distributed selective feedback policy is proposed for the discovery of the optimum relay location with finite average feedback load. It is established that the total number of relays feeding back obeys a Poisson distribution and an analytical expression for the average feedback load is derived. The analytical expressions for the average rate and outage probability with and without selective feedback are also obtained for general path-loss models. It is shown that the optimum location-based relay selection policy outperforms other common relay selection strategies notably. It is also shown that utilizing location information from a small number of relays is enough to achieve almost the same performance with the infinite feedback load case. As generalizations, full-duplex relays, isotropic Poisson point processes, and heterogeneous source-to-relay and relay-to-destination links are also studied.

Index Terms:
Relay selection, cooperative communications, selective feedback, Poisson point processes, stochastic geometry.

I Introduction

I-A Background and Motivation

In wireless networks with large geographical coverage, direct communication links suffer from high propagation losses with distance [3]. To overcome this performance impediment as well as to ensure connectivity in classical and emerging wireless systems, dual-hop communication or cooperative relaying is considered to be an effective transmission and network deployment strategy from a system design point-of-view [4, 5, 6, 7]. At the link level, the classical one-way relay channel supports information flow in one specified direction over potentially shorter transmission distances when compared with direct transmissions. This classical relay channel was originally proposed and studied by van der Meulen in [8]. Cover and El Gamal derived the capacity theorems for Gaussian and certain discrete memoryless relay channels in [9]. They also obtained an achievable lower bound to the capacity of the general relay channel. Subsequently, these results are extended to networks with many relays, multiple antennas (in the form of approximations with bounded capacity gap) and cooperation architectures [10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28].

The notable benefits of relays in wireless communications are increase in network capacity and transmission diversity [16, 17, 18, 19], improved energy efficiency (measured in bits per unit energy) [24, 25, 26] and decrease in network deployment costs [4, 29]. These benefits have already been recognized by the wireless industry, and the possibility of deploying relays for multi-hop communications in emerging wireless systems was included in the latest proposals for LTE-A standards [30, 31]. Two apparent major design problems encountered in industry-grade wireless relay network deployments are, besides link-level capacity optimization, relay selection (due to implementation constraints limiting the number of simultaneous relay connections) and relay placement. Some recent studies showed that significant system-wide performance gains due to optimization over these design degrees-of-freedom can be achieved [32, 33, 34, 35]. However, these aspects are usually eclipsed by the mainstream fixed-relay capacity optimization at the link level, e.g., see the recent surveys [7, 36] and the references therein.

In this paper, we focus on the optimum relay selection problem for randomly deployed single-antenna decode-and-forward (DF) relays connecting source and destination nodes located at arbitrary positions in 2\mathbb{R}^{2}. An example network configuration is illustrated in Fig. 1. In wireless communications, the fading and location processes driving the network capacity and outage events usually vary at different time-scales. For example, the phase of the received electromagnetic waves can change significantly over millisecond intervals, while the magnitude changes are substantial over time intervals in the order of seconds or minutes [3, 37]. By carefully separating the time-scale of changes in fading and location processes, we characterize the key distance balancing and minimum norm (with respect to the mid-point between source and destination nodes) properties for the optimum relay location. We derive the distribution of the channel quality indicator (CQI) at the optimum relay node, which leads to the characterization of the best achievable average rates and the minimum outage probability for the resulting class of two-hop wireless communications paths with optimization over the relay selection dimension. These results hold for general fading distributions and non-increasing path-loss models decaying to zero.

Refer to caption
Figure 1: An example configuration for a relay-aided wireless network.

The time-scale separation approach we adopt for fading and location processes in this paper is motivated by the practical constraints to rule out the prohibitive relay switching rates due to changes in the fading process. In essence, it is similar to the meta-distribution calculations recently introduced for wireless network performance analysis [38, 39, 40]. In most parts of the paper, we focus on a homogeneous Poisson point process (HPPP), which we denote by Φ\Phi, with intensity λ>0\lambda>0 for relay locations. This assumption leads to an insightful analytical structure exposing fundamental dynamics for relay selection. In addition, the uniform relay distribution was also proven to be the optimum configuration for relay placement under some operating regimes such as high signal attenuation as observed in millimeter wave (mmWave) frequency bands [35].111It is well-known that an HPPP can be obtained as the limiting process of a sequence of collection of uniformly distributed points over growing subsets of Euclidean spaces [41]. The generalization of the key results to non-homogeneous PPPs is provided in Section IX, alongside other important extensions such as full-duplex operation (FD) and heterogeneity among the network nodes.

An important aspect of the optimum relay selection problem studied in this paper is the feedback load required to find the solution. As such, the source node (or, a central entity in this regard) requires the knowledge of all relay locations to make the optimum relay selection decision, which presents a major design hurdle for practical network deployments with large numbers of relays. To circumvent this feedback bottleneck, we also develop a low-complexity selective distributed-feedback relay selection policy that requires feedback only from a small number of relays on average.

The proposed feedback mechanism is fully distributed since the relay nodes utilize only their local information to decide whether they feed back or not. The relay selection is still performed centrally but with significantly reduced feedback. We show that the number of relays feeding their CQIs back to the source node for relay selection obeys to a Poisson distribution with a certain mean whose analytical form is completely characterized. We obtain the average rate and outage probability attained by the proposed selective distributed-feedback relay selection policy, and show that it is enough to multiplex only five relays over the feedback channel to achieve almost the same performance as the all-feedback scenario. From an implementation and system-design point of view, this finding presents a massive reduction in feedback load with a negligible loss in communications performance.

I-B Overview of the Main Results and Contributions in Detail

The key parameter to obtain the best achievable rates and minimum outage probability with optimized relay selection is the CQI at the optimum relay node, which we denote by Γopt\Gamma_{\rm opt}. Γopt\Gamma_{\rm opt} is determined by minimization of an appropriately defined relay selection function over all relay locations in Φ\Phi. An important result of this paper is the derivation of the distribution of Γopt\Gamma_{\rm opt}. In particular, its cumulative distribution function (cdf) is shown to be

FΓopt(γ)={0 if γ<d1e2λd2((γd)2arcsec(γd)(γd)21) if γd\begin{split}F_{\Gamma_{\rm opt}}\left(\gamma\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ 1-{\rm e}^{-2\lambda d^{2}\left(\left(\frac{\gamma}{d}\right)^{2}\operatorname{arcsec}\left(\frac{\gamma}{d}\right)-\sqrt{\left(\frac{\gamma}{d}\right)^{2}-1}\right)}&\mbox{ if }\gamma\geq d\end{array}\right.\end{split} (1)

with dd being the half-distance between source and destination nodes. The network performance limits are simply the appropriate integrals of the Φ\Phi-measurable conditional data rates, for which we obtain a single-parameter characterization in the proof of Lemma 1, with respect to FΓopt(γ)F_{\Gamma_{\rm opt}}\left(\gamma\right) on +\mathbb{R}_{+}. As explained in Section VI in detail, the tail distribution of Γopt\Gamma_{\rm opt} decays to zero exponentially and all the moments of Γopt\Gamma_{\rm opt} are finite. Further, it is also shown that Γopt\Gamma_{\rm opt} converges in distribution to dd as λ\lambda\rightarrow\infty.

The previous work on relay selection in random spatial networks is very limited, e.g., see [42, 43, 44, 45, 46, 47, 48, 49, 50, 51]. These papers mostly focus on sub-optimum heuristic strategies for selecting relay nodes such as random and closest-to-source relay selection. The relay nodes selected by the sub-optimum strategies in previous work, as expected, do not possess the fundamental properties of the optimum one that governs the functional form of the distribution of Γopt\Gamma_{\rm opt} given above. These are the distance balancing and minimum norm properties of the optimum relay location. It is currently unknown how close a sub-optimum strategy that does not have these properties to the optimum one. To address this gap in the literature, we obtain a sufficient condition and associated optimality probabilities for a relay selection policy satisfying only one of these fundamental properties to be the optimum selection in Theorems 1, 2 and 3. These results reveal the analytical dependence of relay selection optimality on the network parameters such as relay node intensity, and provide design guidelines for when a sub-optimum policy can be used with small performance loss. To the best of our knowledge, our work in this paper is the first study that rigorously and thoroughly investigates the optimum relay selection problem with randomly deployed relay nodes, and derives the fundamental performance limits for multi-hop relay channels with optimized relay selection.

The selective distributed-feedback relay selection policies with autonomous feedback decision computation at relay nodes are studied in Section VII of the paper. An important result in this part is to show that the total feedback load with threshold-based selection obeys a Poisson distribution with mean μ(T)\mu\left(T\right) given by

μ(T)={0 if T<dλπT22dλT2d22T2λarctan(dT2d2) if Td,\displaystyle\mu\left(T\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }T<d\\ \lambda\pi T^{2}-2d\lambda\sqrt{T^{2}-d^{2}}-2T^{2}\lambda\arctan\left(\frac{d}{\sqrt{T^{2}-d^{2}}}\right)&\mbox{ if }T\geq d\end{array}\right., (4)

where TT is the threshold value with which each relay compares its local CQI to decide to feed back or not. It is seen that μ(T)\mu(T) is a continuous monotone increasing function of TT with μ(d)=0\mu(d)=0 and limTμ(T)=\lim_{T\rightarrow\infty}\mu(T)=\infty. Hence, any given average feedback load can be met by a threshold value by the intermediate value theorem. Further, based on this statistical characterization of the feedback load, we see that the probability of having at least one relay node feeding its CQI back to the source node is equal to 1eμ(T)1-{\rm e}^{-\mu(T)}. Hence, we achieve the same rate and outage performance achieved by the all-feedback strategy with probability at least 0.990.99 if we use a selective distributed-feedback relay selection policy with μ(T)=5\mu(T)=5. The exact characterization of rate and outage with selective feedback is provided in Theorems 6 and 7.

Two significant, as well as surprising, operating regimes for outage probability with selective feedback emerging in Theorem 7 are feedback-limited and rate-limited regimes. To the best of our knowledge, this is the first paper that formally characterizes these operating regimes for the relay selection problem. In particular, the feedback constraints in the feedback-limited operating regime are so tight that we have the same outage probability with or without fading. On the other hand, in the rate-limited operating regime, the target data rate is set so high that we achieve the same outage probability for the all-feedback and selective feedback cases. These operating regimes are discussed in detail in Section VII, with illustrative numerical examples provided in Section VIII.

I-C Paper Organization and Notation

Paper Organization: The rest of the paper is organized as follows. In Section II, we compare and contrast our findings in this paper with relevant previous work in the literature. In Section III, we formally present our system model, the notion of relay selection policy and associated performance metrics to evaluate the relay-aided wireless network performance with optimized relay selection. The optimum relay selection problem is introduced in Section IV. In this section, we also obtain a single-parameter characterization for Φ\Phi-measurable conditional data rates, which underpins most of our analysis by exposing the dependence between the CQI at the selected relay node and the resulting network performance. We discuss the fundamental properties that have to be possessed by the optimum relay node in Section V as well as obtaining a sufficient condition and optimality probabilities for when a relay selection policy possesses only one of these properties.

The best achievable data rates and the minimum outage probability attained by the optimum relay selection policy are derived in Section VI. In this section, we mostly focus on the Rayleigh faded wireless medium for obtaining analytical expressions for network performance, but the same analysis continues to hold for any fading distribution without loss of generality. in Section VII, we put forward the optimum relay selection problem with selective distributed-feedback, obtain a statistical characterization for the total feedback load in the network and derive the resulting network performance with limitations on the average total feedback load. An extensive numerical study is presented in Section VIII to illustrate the main findings of the paper. We provide the extensions of our analysis to full-duplex relays, non-homogeneous PPPs and more heterogeneous communications scenarios with different signal-to-noise ratio (𝖲𝖭𝖱{\sf SNR}) values at relays and the destination node in Section IX. Finally, Section X concludes the paper. Most of our proofs are relegated to Appendices A-I for the sake of exposition.

Notation: The main notation being used throughout the paper is as follows, with some remaining notation introduced later in the paper when needed. We use boldface, upper-case and calligraphic letters to represent vector quantities, random variables and sets, respectively. We use \mathbb{R}, \mathbb{N}, \mathbb{C} and 2\mathbb{R}^{2} to denote the real, natural and complex numbers, and the two-dimensional Euclidean space, respectively. |x||x| denotes the absolute value of a scalar quantity xx (real or complex), whereas 𝒙\|\boldsymbol{x}\| is used to measure the canonical Euclidean norm of a vector quantity 𝒙\boldsymbol{x}. The expected value of a random variable XX is denoted by 𝖤[X]\mathsf{E}\left[X\right]. The probability of an event \mathcal{E} is denoted by 𝖯𝗋()\mathsf{Pr}\left(\mathcal{E}\right). We use 𝟣{}\mathsf{1}_{\left\{\cdot\right\}} to indicate the indicator function of relevant events and mathematical conditions.

II Related Work

Our results and contributions in this paper are related to a large body of work in relay networks. Below, we discuss the most relevant papers under four categories.

II-A Relay Networks Without Location Modeling

The capacity bounds and outage performance of relay networks without explicit modeling of relay locations are well investigated in the literature [29, 52, 53, 54, 55, 56, 57]. In this body of work, the channel states are either fixed and known values, or randomly generated without any modeling of variations in the link signal-to-noise-ratio (𝖲𝖭𝖱{\sf SNR}) as a function of relay locations. The papers such as [29, 52, 53] adopted a communication-theoretic perspective to design single and multiple relay selection strategies, maintaining the full cooperative diversity. The analysis is based on the outage probability and error rate over fading channels. In particular, it was shown in [52], for the case without a direct link over Rayleigh fading channels, that the best-relay selection strategy achieves full diversity with an outage probability asymptotically decaying to zero according to 𝖲𝖭𝖱n\mathsf{SNR}^{-n}, where nn is the number of relays in the network.

Shannon capacity approximations for the Gaussian nn-relay diamond networks were obtained in [54, 55, 56]. The authors in [54] developed a noisy network coding scheme achieving a sum rate within 1.51.5 bits/s/Hz of the cut-set upper bound for the single relay case. In [55], Avestimehr et al. used a deterministic approach to approximate the network capacity of single-relay and two-relay Gaussian diamond networks within 11 bits/s/Hz. The cut-set upper bound was shown to be within 1.96(n+2)1.96(n+2) bits of network capacity by using noisy network coding arguments for the nn-relay case in [56]. The papers [34, 57] developed a network simplification approach by characterising what fraction of total network capacity can be maintained by using only a subset of kk relays, out of nn available relays in the network. It was shown in [34] that every subset of kk relays can at most provide approximately a fraction k/(k+1)k/(k+1) of the total capacity with full-duplex relays. The similar results were extended to the half-duplex case in [57].

Our approach and results differ from this body of work in that we explicitly model the relay locations in the current paper. By doing so, we obtain the best achievable average rates and the minimum outage probability for the resulting class of two-hop wireless communications paths by optimizing over the relay selection dimension. Our results hold for general non-increasing path-loss models decaying to zero, and for general fading distributions in some cases.

II-B Relay Networks With Location Modeling

The outage and rate performance of relay networks with explicit modeling of relay locations are also investigated in the literature [42, 44, 45, 46, 47, 49, 48, 51], usually utilizing sub-optimum techniques for relay selection to have analytical tractability. In this body of work, the relay locations are modeled as points of a spatial point process. The papers such as [45, 49, 48] investigated the relay network performance under the heuristic closest-to-source and mid-point relay selection policies. The authors in [42] considered the relay selection problem from a pre-defined quality-of-service region to minimize the outage probability but they ignored the dependence between source-relay and relay-destination links while calculating the end-to-end network outage performance. In [44], they obtained integral expressions for the outage probability in amplify-and-forward relay networks. The papers [46, 47, 51] considered the relay selection problem to maximize the relay-destination 𝖲𝖭𝖱{\sf SNR} values. This approach is effectively equivalent to the sub-optimum closest-to-destination relay selection policy.

The papers such as [49, 47, 45] considered interference in their relay network model. However, the relay selection policies in these papers do not depend on the network interference level. This is because the network interference introduces a coupling between local relay selection processes at the interfering source-destination links, which makes the optimum multi-user multi-relay selection problem intractable. It also becomes prohibitively hard, even for sub-optimum heuristics taking interference into account for relay selection, to present an accurate statistical characterization of the CQI at the selected relays due to the inter-dependence between local relay selection processes, which does not lead to insightful fundamental expressions for the network performance metrics of interest.

In this paper, different from the previous work on randomly deployed relay networks, we take a more fundamental approach and investigate the performance of the optimum relay selection policy for general path-loss models when there is no direct link between source and destination nodes. We derive structural properties being possessed by the optimum relay selection policy. Moreover, we consider distributed selective feedback relay selection policies, non-homogeneous PPPs and different 𝖲𝖭𝖱{\sf SNR}s at source and relay nodes, which were not considered in the previous work described above. We also consider the time-scale separation between fading and relay location processes cautiously, and identify the effect of such separation on the network performance.

II-C Relay Communications and Feedback

Feedback is an important mechanism in relay networks to facilitate effective relay selection. However, there is only a limited body of work investigating the role of feedback in relay communications [58, 59, 60, 61]. The authors in [58, 59] proposed a compressed sensing approach for joint identity and 𝖲𝖭𝖱{\sf SNR} estimation to select relays with strong channel conditions. Despite significant feedback load reduction, the rates achieved by the developed algorithms in these papers are always bounded away from the full feedback scenario. In [60], the author took a different approach to coordinate asynchronous symbol transmissions from multiple relays by using decision feedback equalizers, shown to achieve full diversity. In [61], the authors developed numerical techniques to solve the optimum power allocation problem, maximizing data rates for single-relay channels, under quantized feedback.

In this paper, we focus on solving the feedback problem for relays with random locations by proposing a threshold-based distributed selective feedback relay selection policy to reduce the network feedback load. To the best of our knowledge, there is no prior work addressing this problem when relays are randomly distributed, where availability of relay location and/or channel state information at a central entity or the source node is a common assumption to perform relay selection. This operating assumption requires an excessive feedback load, which does not scale well for large networks. We show that it is enough to multiplex only five relays over the feedback channel to achieve almost the same rate performance as the all-feedback scenario. This is an important reduction in the feedback load from a system-design point of view.

II-D Routing in Spatial Networks

There is a large and growing body of work on routing in random spatial networks [62, 63, 64, 65, 66, 67, 68, 69, 70], to which our results in this paper are also related. This body of work focuses on the evaluation of network layer performance metrics such as end-to-end delay and throughput to assess the efficiency of a given routing protocol. The analysis is usually carried out for fixed rate transmissions with threshold-based detection and ALOHA- or TDMA-type MAC layer. Some models are very specific, focusing on line networks such as [64, 68, 69, 70]. In contrast, we take a more fundamental approach in the current paper by considering the information capacity of cooperative relay channels as the performance metric of interest. This change in the approach also triggers a change in the mathematical analysis, different than the above previous work.

The second fundamental difference between our paper and the previous routing work is that we maximize the “routing” performance of two-hop relay networks by optimizing over the relay selection dimension. The previous routing work fixes the routing protocol such as the kk-nearest neighbor routing or nearest neighbor routing with minimum bounded progress (i.e., called quasi-equal distance (QED) routing in [70]). To the best of our knowledge, there is no prior paper in the routing literature optimizing the selection of intermediate nodes (from a random spatial field) to maximize the network performance, as we analyzed thoroughly in the current paper.

III Network Model and Performance Metrics

III-A Network Model

We consider a relay-aided spatial wireless network in 2\mathbb{R}^{2}, as illustrated by Fig. 1. The network contains a source-destination pair having arbitrary locations 𝒙s2\boldsymbol{x}_{\rm s}\in\mathbb{R}^{2} (source node) and 𝒙d2\boldsymbol{x}_{\rm d}\in\mathbb{R}^{2} (destination node). The locations of potential DF relay nodes in the network are given by φ={𝒙1,𝒙2,}\varphi=\left\{\boldsymbol{x}_{1},\boldsymbol{x}_{2},\ldots\right\}, where 𝒙i2\boldsymbol{x}_{i}\in\mathbb{R}^{2} represents the iith DF relay location for ii\in\mathbb{N}. We will always assume that φ\varphi is a locally finite set, i.e., there are only finitely many relays in every bounded subset of 2\mathbb{R}^{2}. For performance analysis, we will consider the communication scenario in which relay locations are random and determined according to a spatial homogeneous Poisson point process (HPPP) Φ={𝑿1,𝑿2,}\Phi=\left\{\boldsymbol{X}_{1},\boldsymbol{X}_{2},\ldots\right\}. Hence, φ\varphi should be interpreted as a particular realization of Φ\Phi below, which will usually be clear from the context unless otherwise stated. Relay-based cellular networks are an important example of this model [71, 72, 4], although our analysis is not restricted to infrastructure-based wireless communications. We assume that the primary aim of the relays is to assist data communication between source and destination nodes without generating any additional traffic, as in the latest proposals for LTE-A standards [30, 31].

For the sake of exposition, we will focus our attention only on the easy-to-implement half-duplex (HD) relaying operation in the remainder of the paper. As explained in Section IX, this assumption will not limit our results in the full-duplex (FD) case [19]. The HD operation in our setup follows the classical resource orthogonalization approach [16], and we consider that half of the degrees-of-freedom available in the wireless channel is used by the source node, while the other half is allocated for the relay-destination link. This is usually done through time division multiplexing (TDM) in existing wireless systems when relay and source nodes share the same frequency band [72]. In a relay-aided wireless network setting, the location of the selected relay node is of critical importance to determine the achievable data rates between source and destination. To this end, we examine the network performance under a given relay selection policy, which is formally defined as follows.

Definition 1

A relay selection policy 𝒫:Σ2\mathcal{P}:\Sigma\mapsto\mathbb{R}^{2} is a mapping from the set of all finite subsets of 2\mathbb{R}^{2}, denoted by Σ\Sigma, to 2\mathbb{R}^{2} satisfying the condition 𝒫(φ)φ\mathcal{P}\left(\varphi\right)\in\varphi for all φΣ\varphi\in\Sigma.

Some examples of 𝒫\mathcal{P} include the ones choosing the relay closest to the source, the relay closest to the destination and the relay closest to the mid-point between source and destination. Our focus below is predominantly on the statistical characterization of the performance of a given relay selection policy 𝒫\mathcal{P} over a random ensemble of relay locations. For the sake of notational simplicity, we parametrize the relay location selected by 𝒫\mathcal{P} as 𝒙𝒫\boldsymbol{x}_{\mathcal{P}} in the remainder of the paper, with the understanding that 𝒙𝒫=𝒫(φ)\boldsymbol{x}_{\mathcal{P}}=\mathcal{P}\left(\varphi\right) for any given φΣ\varphi\in\Sigma. The same meaning is attributed to other variables throughout the paper when we use this parametrization for them as well. The instantaneous rate of information flow from source to destination (in bits/s/Hz) as a function of the relay selection policy 𝒫\mathcal{P} is given as

𝗋φ(𝒫)\displaystyle{\sf r}_{\varphi}\left(\mathcal{P}\right) =\displaystyle= 12min{log2(1+𝖲𝖭𝖱|Hs,r|2G(𝒙s𝒙𝒫)),\displaystyle\frac{1}{2}\min\bigl{\{}\log_{2}\left(1+{\sf SNR}\left|H_{\rm s,r}\right|^{2}G\left(\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}_{\mathcal{P}}\right\|\right)\right), (5)
log2(1+𝖲𝖭𝖱(|Hs,d|2G(𝒙s𝒙d)+|Hr,d|2G(𝒙𝒫𝒙d)))}\displaystyle\quad\log_{2}\left(1+{\sf SNR}\left(\left|H_{\rm s,d}\right|^{2}G\left(\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}_{\rm d}\right\|\right)+\left|H_{\rm r,d}\right|^{2}G\left(\left\|\boldsymbol{x}_{\mathcal{P}}-\boldsymbol{x}_{\rm d}\right\|\right)\right)\right)\bigr{\}}

where GG is a non-negative non-increasing path-loss function decaying to zero, Ha,bH_{a,b}\in\mathbb{C} is the random fading coefficient between the source, selected relay and destination nodes for a{s,r}a\in\left\{{\rm s,r}\right\} and b{r,d}b\in\left\{{\rm r,d}\right\}, and 𝖲𝖭𝖱PWN0{\sf SNR}\triangleq\frac{P}{WN_{0}} is the signal-to-noise ratio with PP, WW and N0N_{0} being transmission power, communication bandwidth and noise energy per complex dimension, respectively, [37]. The rate 𝗋φ(𝒫){\sf r}_{\varphi}\left(\mathcal{P}\right) in (5) is achievable for each fading state 𝑯=(Hs,r,Hr,d,Hs,d)\boldsymbol{H}=\left(H_{\rm s,r},H_{\rm r,d},H_{\rm s,d}\right)^{\top} by using independent Gaussian codebooks at the source and relay nodes and channel state information (CSI) at the receivers [16].222Implicit in this formulation is the assumption of a perfect multiple-access and interference cancellation scheme so that the background noise is the only additive channel distortion at the receivers.

One simplifying assumption that we will make for the rate function 𝗋φ(𝒫){\sf r}_{\varphi}\left(\mathcal{P}\right) in order to pose the optimum relay selection problem for general path-loss models is as follows. We will assume that the signal power received over the longer source-destination link is much smaller than the one received over the shorter relay-destination link. This assumption corresponds to the physical circumstances in which either the direct link between source and destination nodes is severely shadowed by an object in the environment (i.e., |Hs,d|0\left|H_{\rm s,d}\right|\approx 0), or the decay of GG with distance is very sharp as in the mmWave communications [72, 71, 4]. It models the multi-hop mode of operation for relay channels [19] as well as the common LTE-A relay deployment scenarios such as dead spot elimination, coverage extension to rural areas and emergency coverage [72, 71, 4].

We will also assume that random fading coefficients Hs,rH_{\rm s,r} and Hr,dH_{\rm r,d} change at a much faster time-scale than the network node locations, which is usually the case in typical wireless communication scenarios [3, 73]. In such cases, it is an onerous task, if not practical due to the triggered excessive relay switching rate, for the source node to obtain CSI for all source-to-relay and relay-to-destination channels, and establish a connection to another relay node for handing over the data traffic each time fading coefficients change. Hence, our selection criterion will be based only on relay locations for selecting the relay node, which is also embodied in Definition 1. This approach is reminiscent of the base station selection strategy in cellular networks. This is also the optimum approach when only location information but not the full CSI is available at the source node.333We continue to assume the availability of full CSI at the receiver side of each link so that the data rates in (5) are achievable.

III-B Performance Metrics

For determining the performance of a relay selection policy 𝒫\mathcal{P}, we will use the outage probability and average data rate as our performance metrics [3]. The former one is the common metric to measure the system performance for delay-sensitive data traffic requiring a minimum data rate for successful communications, whereas the latter is more appropriate for when the delay requirement is not stringent and transmitters are allowed to adjust their transmission rates based on the observed path-loss values [74]. We assume for both cases that the permissible decoding delay is large enough to average over the fading process. Then, given 𝒫\mathcal{P} and the random relay locations Φ={𝑿1,𝑿2,}\Phi=\left\{\boldsymbol{X}_{1},\boldsymbol{X}_{2},\ldots\right\}, the achievable rate over the wireless link connecting the source and selected relay is 𝖤[log2(1+𝖲𝖭𝖱|Hs,r|2G(𝒙s𝑿𝒫))|Φ]\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H_{\rm s,r}\right|^{2}G\left(\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{X}_{\mathcal{P}}\right\|\right)\right)\big{|}\Phi\right], where the expectation is taken only over the randomness due to fading. Similarly, the achievable rate between the selected relay and destination is 𝖤[log2(1+𝖲𝖭𝖱|Hr,d|2G(𝑿𝒫𝒙d))|Φ]\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H_{\rm r,d}\right|^{2}G\left(\left\|\boldsymbol{X}_{\mathcal{P}}-\boldsymbol{x}_{\rm d}\right\|\right)\right)\big{|}\Phi\right].444For the rate between the selected relay and destination, we ignore the term containing |Hs,d|\left|H_{\rm s,d}\right| since we assumed |Hs,d|0\left|H_{\rm s,d}\right|\approx 0. If this is not an appropriate modeling assumption, then the relay-destination rate should be interpreted as the rate obtained through a type of selection combining in which the destination takes only the signals from the relay into account for data decoding. If maximum ratio combining is employed at the destination, similar performance analysis continues to hold as discussed in Section VIII. However, we cannot claim the optimality of the proposed relay selection policy for general path-loss models in this case. Hence, we can express the data rate, averaged over the fading process, from source to destination as

𝖱Φ(𝒫)=12min{𝖤[log2(1+𝖲𝖭𝖱|Hs,r|2G(𝒙s𝑿𝒫))|Φ],\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}\right)=\frac{1}{2}\min\bigl{\{}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H_{\rm s,r}\right|^{2}G\left(\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{X}_{\mathcal{P}}\right\|\right)\right)\big{|}\Phi\right],
𝖤[log2(1+𝖲𝖭𝖱|Hr,d|2G(𝑿𝒫𝒙d))|Φ]},\displaystyle\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H_{\rm r,d}\right|^{2}G\left(\left\|\boldsymbol{X}_{\mathcal{P}}-\boldsymbol{x}_{\rm d}\right\|\right)\right)\big{|}\Phi\right]\bigr{\}}, (6)

which is a function of 𝒫\mathcal{P} and Φ\Phi. Using 𝖱Φ(𝒫){\sf R}_{\Phi}\left(\mathcal{P}\right) in (6), the outage probability for delay-sensitive data traffic and the average rate for elastic data traffic for which the variable rate transmission is permissible (e.g., dynamic adaptive video streaming over HTTP [75]) as metrics indicative of the relay-assisted network performance are defined as below.

Definition 2

For a target bit rate ρ\rho and a relay selection policy 𝒫\mathcal{P}, the outage probability 𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}\right) is equal to

𝖯out(𝒫)=𝖯𝗋{𝖱Φ(𝒫)ρ}.\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}\right)=\mathsf{Pr}\left\{{\sf R}_{\Phi}\left(\mathcal{P}\right)\leq\rho\right\}. (7)
Definition 3

For a given relay selection policy 𝒫\mathcal{P}, the average rate is defined to be

𝖱ave(𝒫)=𝖤[𝖱Φ(𝒫)],\displaystyle{\sf R}_{\rm ave}\left(\mathcal{P}\right)=\mathsf{E}\left[{\sf R}_{\Phi}\left(\mathcal{P}\right)\right], (8)

where the expectation is over the random relay locations.

It is important to note that 𝖱ave(𝒫){\sf R}_{\rm ave}\left(\mathcal{P}\right) defined in (8) should be interpreted as the average of the rate in (6), which is averaged over the random relay locations. It is not the ergodic rate achieved over the long time-horizon for the same source, relay and destination nodes. Rather, the rate in (6) is attained with a different relay node selected by 𝒫\mathcal{P} for each realization of Φ\Phi. In addition, we observe that we can change the order of minimum and expectation operators while going from (5) to (6). This is because the decoding delays permit averaging over the fading process before switching to another relay node. It is more advantageous to minimize expected rates than averaging the minimum of instantaneous rates. This is the waiting-gain we have in (6).

IV Optimum Relay Selection Problem

Consider the relay selection function s^(𝒙)\widehat{s}\left(\boldsymbol{x}\right), which is defined as

s^(𝒙)max{𝒙s𝒙,𝒙𝒙d}\displaystyle\widehat{s}\left(\boldsymbol{x}\right)\triangleq\max\left\{\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}\right\|,\left\|\boldsymbol{x}-\boldsymbol{x}_{\rm d}\right\|\right\} (9)

for 𝒙2\boldsymbol{x}\in\mathbb{R}^{2}. We say s^(𝒙)\widehat{s}\left(\boldsymbol{x}\right) is the CQI indicator of the relay located at 𝒙2\boldsymbol{x}\in\mathbb{R}^{2}. For a given set of relay locations φΣ\varphi\in\Sigma, we will start with minimization of s^(𝒙)\widehat{s}\left(\boldsymbol{x}\right) over φ\varphi. The solution of this optimization problem is closely related to optimizing the performance metrics 𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}\right) and 𝖱ave(𝒫){\sf R}_{\rm ave}\left(\mathcal{P}\right), as we will formally establish below.

We write the optimum relay selection problem as

minimize𝒙2s^(𝒙)subject to𝒙φ\displaystyle\begin{array}[]{ll}\underset{\boldsymbol{x}\in\mathbb{R}^{2}}{\mbox{minimize}}&\widehat{s}\left(\boldsymbol{x}\right)\\ \mbox{subject to}&\boldsymbol{x}\in\varphi\end{array} (12)

for each φΣ\varphi\in\Sigma. The solution of (12) is well-defined and belongs to φ\varphi for all φΣ\varphi\in\Sigma since φ\varphi is assumed to be a locally finite set. The optimum relay selection policy, which we denote by 𝒫opt\mathcal{P}_{\rm opt}, is the one that solves (12) for all φΣ\varphi\in\Sigma. When we write 𝒙opt\boldsymbol{x}_{\rm opt}, we will refer to the location of the relay selected by 𝒫opt\mathcal{P}_{\rm opt}. Observing the structure of s^(𝒙)\widehat{s}\left(\boldsymbol{x}\right) and min-max type of optimization in (12), it can be seen that the optimum relay location in φ\varphi must have two important properties: (i) distance balancing property (with respect to the source and destination locations) and (ii) minimum norm property (with respect to the mid-point between source and destination). We will investigate these properties in detail in Section V.

The connection between the solution of the optimization problem in (12) and the performance metrics 𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}\right) and 𝖱ave(𝒫){\sf R}_{\rm ave}\left(\mathcal{P}\right) is not explicit. In Lemma 1 below, we relate the solution of (12) to 𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}\right) and 𝖱ave(𝒫){\sf R}_{\rm ave}\left(\mathcal{P}\right) by showing that the optimum relay selection policy solving (12) for each φΣ\varphi\in\Sigma is also the best choice for optimizing 𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}\right) and 𝖱ave(𝒫){\sf R}_{\rm ave}\left(\mathcal{P}\right).

Lemma 1

Let Ξ\Xi be the set of all relay selection policies. Then, for identically distributed fading at each wireless link, we have

𝖯out(𝒫opt)=inf𝒫Ξ𝖯out(𝒫)\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right)=\inf_{\mathcal{P}\in\Xi}{\sf P}_{\rm out}\left(\mathcal{P}\right) (13)
𝖱ave(𝒫opt)=sup𝒫Ξ𝖱ave(𝒫).\displaystyle{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right)=\sup_{\mathcal{P}\in\Xi}{\sf R}_{\rm ave}\left(\mathcal{P}\right). (14)
Proof:

We first observe the following trivial inequality. If X1X_{1} and X2X_{2} are two identically distributed non-negative random variables, and c1c_{1} and c2c_{2} are two non-negative arbitrary constants satisfying c1c2c_{1}\leq c_{2}, then 𝖤[log2(1+c1X1)]𝖤[log2(1+c2X2)]\mathsf{E}\left[\log_{2}\left(1+c_{1}X_{1}\right)\right]\leq\mathsf{E}\left[\log_{2}\left(1+c_{2}X_{2}\right)\right]. Using this inequality and recalling that for any given relay selection policy 𝒫\mathcal{P}, 𝑿𝒫\boldsymbol{X}_{\mathcal{P}} is the location of the relay node selected by 𝒫\mathcal{P} when it runs over Φ\Phi, we can write 𝖱Φ(𝒫)=12𝖤[log2(1+𝖲𝖭𝖱|H|2G(s^(𝑿𝒫)))|Φ]{\sf R}_{\Phi}\left(\mathcal{P}\right)=\frac{1}{2}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H\right|^{2}G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)\right)\big{|}\Phi\right] where HH is the generic fading random variable having the same distribution with Hs,rH_{\rm s,r} and Hr,dH_{\rm r,d}. Since 𝒫opt\mathcal{P}_{\rm opt} solves (12) for each φΣ\varphi\in\Sigma, we also have s^(𝑿opt)s^(𝑿𝒫)\widehat{s}\left(\boldsymbol{X}_{{\rm opt}}\right)\leq\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right). Using the above inequality one more time and the property that GG is a non-increasing function, we conclude that 𝖱Φ(𝒫opt)𝖱Φ(𝒫){\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right)\geq{\sf R}_{\Phi}\left(\mathcal{P}\right) for any 𝒫Ξ\mathcal{P}\in\Xi. This implies 𝖱ave(𝒫opt)=sup𝒫Ξ𝖱ave(𝒫){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right)=\sup_{\mathcal{P}\in\Xi}{\sf R}_{\rm ave}\left(\mathcal{P}\right) and 𝖯out(𝒫opt)=inf𝒫Ξ𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right)=\inf_{\mathcal{P}\in\Xi}{\sf P}_{\rm out}\left(\mathcal{P}\right). ∎

An important remark about the proof of Lemma 1 is that it also shows 𝖱Φ(𝒫){\sf R}_{\Phi}\left(\mathcal{P}\right) can be written as

𝖱Φ(𝒫)=12𝖤[log2(1+𝖲𝖭𝖱|H|2G(s^(𝑿𝒫)))|Φ]\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}\right)=\frac{1}{2}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H\right|^{2}G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)\right)\big{|}\Phi\right] (15)

for any relay selection policy 𝒫\mathcal{P}. Since 𝑿𝒫\boldsymbol{X}_{\mathcal{P}} is a Φ\Phi-measurable random variable and HH is independent of Φ\Phi, 𝖱Φ(𝒫){\sf R}_{\Phi}\left(\mathcal{P}\right) is as well equal to

𝖱Φ(𝒫)=12𝖤[log2(1+𝖲𝖭𝖱|H|2G(s^(𝑿𝒫)))|𝑿𝒫].\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}\right)=\frac{1}{2}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H\right|^{2}G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)\right)\big{|}\boldsymbol{X}_{\mathcal{P}}\right]. (16)

Therefore, in order to obtain 𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}\right) or 𝖱ave(𝒫){\sf R}_{\rm ave}\left(\mathcal{P}\right), we first need to obtain the distribution of 𝑿𝒫\boldsymbol{X}_{\mathcal{P}} or that of s^(𝑿𝒫)\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right), and calculate relevant averages with respect to these distributions. For 𝒫opt\mathcal{P}_{\rm opt} with HPPP distributed relays, the distribution of s^(𝑿opt)\widehat{s}\left(\boldsymbol{X}_{\rm opt}\right) is not known and it will be derived in Section VI to obtain 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) and 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right).

V Key Properties of the Optimum Policy

As articulated in Section IV, the optimum relay selection policy 𝒫opt\mathcal{P}_{\rm opt} solving (12) must possess two key properties of distance balancing and minimum norm. In this section, we will characterize these properties in detail before we establish the statistical structure of s^(𝑿opt)\widehat{s}\left(\boldsymbol{X}_{\rm opt}\right) for Poisson distributed relays over the plane in Section VI. In order to put our discussion into perspective, we will make use of the heuristic mid-point relay selection policy 𝒫mid\mathcal{P}_{\rm mid}, which selects the relay having the closest location 𝑿mid\boldsymbol{X}_{\rm mid} to the mid-point between source and destination. 𝒫mid\mathcal{P}_{\rm mid} only possesses one of these key properties, which is the minimum norm property. We will obtain the probability of 𝒫mid\mathcal{P}_{\rm mid} being optimum for Poisson distributed relays as well as a sufficient condition that guarantees its optimality. This analysis will reveal the importance of possessing both properties for optimality and the fact that probability of 𝒫mid\mathcal{P}_{\rm mid} being optimum is small for large relay intensity and separation between source and destination nodes. This finding will further motivate our analysis in Section VI.

V-A Minimum Norm and Distance Balancing Properties

We will start with a set of arbitrary relay locations φΣ\varphi\in\Sigma without imposing any statistical structure, and then analyze the case in which relay nodes are randomly distributed over the plane according to an HPPP with intensity λ>0\lambda>0. Below, 𝒘\boldsymbol{w} will denote the mid-point between source and destination nodes, i.e., 𝒘=𝒙s+𝒙d2\boldsymbol{w}=\frac{\boldsymbol{x}_{\rm s}+\boldsymbol{x}_{\rm d}}{2}, and dd will denote the distance from 𝒘\boldsymbol{w} to 𝒙s\boldsymbol{x}_{\rm s} or to 𝒙d\boldsymbol{x}_{\rm d}, i.e., d=𝒙s𝒘=𝒙d𝒘d=\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{w}\right\|=\left\|\boldsymbol{x}_{\rm d}-\boldsymbol{w}\right\|. The following simple result formally states the minimum norm property for optimum relay locations. This result also provides a motivation for the mid-point relay selection rule.

Lemma 2

For all 𝐱2\boldsymbol{x}\in\mathbb{R}^{2}, s^(𝐰)s^(𝐱)\widehat{s}\left(\boldsymbol{w}\right)\leq\widehat{s}\left(\boldsymbol{x}\right).

Proof:

By triangle inequality, we have 𝒙s𝒙+𝒙𝒙d2d\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}\right\|+\left\|\boldsymbol{x}-\boldsymbol{x}_{\rm d}\right\|\geq 2d for all 𝒙2\boldsymbol{x}\in\mathbb{R}^{2}, which implies s^(𝒙)d\widehat{s}\left(\boldsymbol{x}\right)\geq d for all 𝒙2\boldsymbol{x}\in\mathbb{R}^{2}. This lower bound is achieved with equality when 𝒙=𝒘\boldsymbol{x}=\boldsymbol{w}. ∎

Lemma 2 suggests that 𝒫mid\mathcal{P}_{\rm mid} possesses one of the key properties for being a solution for (12) since it always chooses the relay node closest to 𝒘\boldsymbol{w}, which is globally the best location to place a relay node without direct link signal reception. However, 𝒫mid\mathcal{P}_{\rm mid} does not result in the optimum relay selection for all relay configurations φΣ\varphi\in\Sigma since, in addition to its distance from the mid-point between 𝒙s\boldsymbol{x}_{\rm s} and 𝒙d\boldsymbol{x}_{\rm d}, the relay’s orientation is also crucial in determining the value of s^(𝒙)\widehat{s}\left(\boldsymbol{x}\right). That is, the closer the relay location 𝒙\boldsymbol{x} to the equidistant hyperplane \mathcal{H} between 𝒙s\boldsymbol{x}_{\rm s} and 𝒙d\boldsymbol{x}_{\rm d} is, it better balances the distances between source and destination and leads to a smaller value that s^(𝒙)\widehat{s}\left(\boldsymbol{x}\right) takes. We will utilize the distance balancing property for optimum relay locations to obtain a sufficient condition under which 𝒫mid\mathcal{P}_{\rm mid} is optimum. We formally state this property in the following lemma.

Lemma 3

Let ={𝐱2:(𝐱s𝐱d)𝐱=12(𝐱s2𝐱d2)}\mathcal{H}=\left\{\boldsymbol{x}\in\mathbb{R}^{2}:\left(\boldsymbol{x}_{\rm s}-\boldsymbol{x}_{\rm d}\right)^{\top}\boldsymbol{x}=\frac{1}{2}\left(\left\|\boldsymbol{x}_{\rm s}\right\|^{2}-\left\|\boldsymbol{x}_{\rm d}\right\|^{2}\right)\right\} be the equidistant hyperplane between 𝐱s\boldsymbol{x}_{\rm s} and 𝐱d\boldsymbol{x}_{\rm d}. For a given 𝐲2\boldsymbol{y}\in\mathbb{R}^{2}, let also 𝒞={𝐱2:𝐱𝐰=𝐲𝐰}\mathcal{C}=\left\{\boldsymbol{x}\in\mathbb{R}^{2}:\left\|\boldsymbol{x}-\boldsymbol{w}\right\|=\left\|\boldsymbol{y}-\boldsymbol{w}\right\|\right\}, which is the circle around 𝐰\boldsymbol{w} with radius 𝐲𝐰\left\|\boldsymbol{y}-\boldsymbol{w}\right\|. Then, for any 𝐱𝒞\boldsymbol{x}\in\mathcal{H}\cap\mathcal{C}, we have s^(𝐱)=d2+𝐲𝐰2\widehat{s}\left(\boldsymbol{x}\right)=\sqrt{d^{2}+\left\|\boldsymbol{y}-\boldsymbol{w}\right\|^{2}} and s^(𝐱)s^(𝐲)\widehat{s}\left(\boldsymbol{x}\right)\leq\widehat{s}\left(\boldsymbol{y}\right).

Proof:

Assume 𝒚\boldsymbol{y} belongs to the half-space closer to 𝒙d\boldsymbol{x}_{\rm d} and consider the decomposition 𝒚=𝒘+𝒛\boldsymbol{y}=\boldsymbol{w}+\boldsymbol{z} for some 𝒛2\boldsymbol{z}\in\mathbb{R}^{2}. Then, we have (𝒙s𝒙d)𝒛0\left(\boldsymbol{x}_{\rm s}-\boldsymbol{x}_{\rm d}\right)^{\top}\boldsymbol{z}\leq 0 and

s^(𝒚)\displaystyle\widehat{s}\left(\boldsymbol{y}\right) =𝒙s𝒚=𝒙s𝒘2(𝒙s𝒙d)𝒛+𝒛2𝒙s𝒘2+𝒚𝒘2.\displaystyle=\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{y}\right\|=\sqrt{\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{w}\right\|^{2}-\left(\boldsymbol{x}_{\rm s}-\boldsymbol{x}_{\rm d}\right)^{\top}\boldsymbol{z}+\left\|\boldsymbol{z}\right\|^{2}}\geq\sqrt{\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{w}\right\|^{2}+\left\|\boldsymbol{y}-\boldsymbol{w}\right\|^{2}}.

Similarly, for any 𝒙𝒞\boldsymbol{x}\in\mathcal{H}\cap\mathcal{C}, we have s^(𝒙)=𝒙s𝒘2+𝒚𝒘2\widehat{s}\left(\boldsymbol{x}\right)=\sqrt{\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{w}\right\|^{2}+\left\|\boldsymbol{y}-\boldsymbol{w}\right\|^{2}}. Hence, s^(𝒙)s^(𝒚)\widehat{s}\left(\boldsymbol{x}\right)\leq\widehat{s}\left(\boldsymbol{y}\right). The arguments for when 𝒚\boldsymbol{y} belongs to the half-space closer to 𝒙s\boldsymbol{x}_{\rm s} are the same. ∎

V-B Optimality Probability for 𝒫mid\mathcal{P}_{\rm mid}

Using Lemma 3, we obtain a sufficient condition for the optimality of 𝒫mid\mathcal{P}_{\rm mid} in the next theorem.

Theorem 1

Let 𝐱midφ\boldsymbol{x}_{\rm mid}\in\varphi be the relay location selected by 𝒫mid\mathcal{P}_{\rm mid} and 𝐱(2)φ\boldsymbol{x}_{(2)}\in\varphi be the location of the second closest relay to 𝐰\boldsymbol{w}. Then, 𝒫mid\mathcal{P}_{\rm mid} solves (12) if s^(𝐱mid)d2+𝐱(2)𝐰2\widehat{s}\left(\boldsymbol{x}_{\rm mid}\right)\leq\sqrt{d^{2}+\left\|\boldsymbol{x}_{(2)}-\boldsymbol{w}\right\|^{2}}.

Proof:

Let 𝒙(1),𝒙(2),\boldsymbol{x}_{(1)},\boldsymbol{x}_{(2)},\ldots be the ordering of relay locations in φ\varphi in an ascending manner with respect to their distances to 𝒘\boldsymbol{w}, i.e., 𝒙mid=𝒙(1)\boldsymbol{x}_{\rm mid}=\boldsymbol{x}_{(1)} and 𝒙(i)\boldsymbol{x}_{(i)} is the location of the iith closest relay to 𝒘\boldsymbol{w}. Then, by Lemma 3, d2+𝒙(i)𝒘2s^(𝒙(i))\sqrt{d^{2}+\left\|\boldsymbol{x}_{(i)}-\boldsymbol{w}\right\|^{2}}\leq\widehat{s}\left(\boldsymbol{x}_{(i)}\right) for all i1i\geq 1. Further, d2+𝒙(2)𝒘2d2+𝒙(i)𝒘2\sqrt{d^{2}+\left\|\boldsymbol{x}_{(2)}-\boldsymbol{w}\right\|^{2}}\leq\sqrt{d^{2}+\left\|\boldsymbol{x}_{(i)}-\boldsymbol{w}\right\|^{2}} for all i2i\geq 2. Hence, whenever s^(𝒙mid)d2+𝒙(2)𝒘2\widehat{s}\left(\boldsymbol{x}_{\rm mid}\right)\leq\sqrt{d^{2}+\left\|\boldsymbol{x}_{(2)}-\boldsymbol{w}\right\|^{2}}, we have s^(𝒙mid)=min𝒙φs^(𝒙)\widehat{s}\left(\boldsymbol{x}_{\rm mid}\right)=\min_{\boldsymbol{x}\in\varphi}\widehat{s}\left(\boldsymbol{x}\right). ∎

For any realization of relay locations, this condition is easy to check since we only need to know the location information of the closest and second closest relay nodes to the mid-point. Now, by considering random relay locations given according to Φ\Phi, which is an HPPP having intensity λ>0\lambda>0 (i.e., λ\lambda is the average number of relays per unit area), we obtain the probability with which the sufficient condition in Theorem 1 is satisfied.

Theorem 2

Let Φ\Phi be an HPPP having intensity λ>0\lambda>0. Let suff\mathcal{E}_{\rm suff} be the event for the sufficient condition given in Theorem 1 for the optimality of 𝒫mid\mathcal{P}_{\rm mid}. Then, 𝖯𝗋(suff)\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) is equal to

𝖯𝗋(suff)=eλπd2erfc(λπd).\displaystyle\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right)={\rm e}^{\lambda\pi d^{2}}\text{erfc}\left(\sqrt{\lambda\pi}\,d\right). (17)
Proof:

See Appendix A. ∎

In the next theorem, we extend the result in Theorem 2 and provide an expression for the probability that 𝒫mid\mathcal{P}_{\rm mid} is optimum for Poisson distributed relay locations. Even though it will be more complicated than the expression for 𝖯𝗋(suff)\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) given in (17), it is the exact expression for the probability 𝖯𝗋{𝑿mid=𝑿opt}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\}, where 𝑿midΦ\boldsymbol{X}_{\rm mid}\in\Phi represents the random location of the relay chosen by 𝒫mid\mathcal{P}_{\rm mid}.

Theorem 3

Let Φ\Phi be an HPPP having intensity λ>0\lambda>0. For 𝐱2\boldsymbol{x}\in\mathbb{R}^{2} with norm ψ\psi and angle θ[0,π2]\theta\in\left[0,\frac{\pi}{2}\right], let P(𝐱)P\left(\boldsymbol{x}\right) be the function defined as

P(𝒙)=((s^(𝒙))2ψ2)(π2θ)d2sin(2θ)V(𝒙,d)+V(𝒙,dsin(θ)),\displaystyle P\left(\boldsymbol{x}\right)=\left(\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}-\psi^{2}\right)\left(\pi-2\theta\right)-d^{2}\sin\left(2\theta\right)-V\left(\boldsymbol{x},d\right)+V\left(\boldsymbol{x},d\sin\left(\theta\right)\right),

where V(𝐱,y)=2y(s^(𝐱))2y2+2(s^(𝐱))2arctan(y(s^(𝐱))2y2)V\left(\boldsymbol{x},y\right)=2y\sqrt{\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}-y^{2}}+2\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}\arctan\left(\frac{y}{\sqrt{\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}-y^{2}}}\right). Then, 𝖯𝗋{𝐗mid=𝐗opt}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} is equal to

𝖯𝗋{𝑿mid=𝑿opt}=4𝖤[exp(λP(𝑿mid))𝟣{0Θmidπ2}],\displaystyle\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\}=4\mathsf{E}\left[\exp\left(-\lambda P\left(\boldsymbol{X}_{\rm mid}\right)\right)\mathsf{1}_{\left\{0\leq\Theta_{\rm mid}\leq\frac{\pi}{2}\right\}}\right], (18)

where 𝟣{}\mathsf{1}_{\left\{\cdot\right\}} is the indicator function and Θmid\Theta_{\rm mid} is the angle of 𝐗mid\boldsymbol{X}_{\rm mid}.

Proof:

See Appendix C. ∎

We note that it is numerically easy to calculate the expectation in (18) by using the probability density function (pdf) of 𝑿mid\boldsymbol{X}_{\rm mid}. In particular, Θmid\Theta_{\rm mid} is uniformly distributed over [0,2π)\left.\left[0,2\pi\right)\right., Ψmid=𝑿mid\Psi_{\rm mid}=\left\|\boldsymbol{X}_{\rm mid}\right\| has the nearest-neighbor pdf fNN(ψ)=2λπψeλπψ2𝟣{ψ0}f_{\rm NN}\left(\psi\right)=2\lambda\pi\psi{\rm e}^{-\lambda\pi\psi^{2}}\mathsf{1}_{\left\{\psi\geq 0\right\}}, and Θmid\Theta_{\rm mid} and Ψmid\Psi_{\rm mid} are independent random variables.

Refer to captionRefer to caption
Figure 2: 𝖯𝗋(suff)\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) and 𝖯𝗋{𝑿mid=𝑿opt}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} as a function of λ\lambda for various values of dd.

Fig. 2 demonstrates the analytical expressions derived for 𝖯𝗋(suff)\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) and 𝖯𝗋{𝑿mid=𝑿opt}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} in Theorem 2 and Theorem 3 as well as the simulated probability values. Our results in these theorems are correct for all source and destination locations, however we will assume 𝒙s=(d,0)\boldsymbol{x}_{\rm s}=\left(-d,0\right) and 𝒙d=(d,0)\boldsymbol{x}_{\rm d}=\left(d,0\right) to explain our main observations in Fig. 2 without loss of generality. As can be observed in this figure, 𝖯𝗋{𝑿mid=𝑿opt}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} decreases as the relay intensity or the separation between source and destination nodes increases. The diminishing behaviour of 𝖯𝗋{𝑿mid=𝑿opt}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} as a function of relay intensity arises from the fact that it becomes more likely to find a relay node 𝑿Φ\boldsymbol{X}\in\Phi with a better distance balancing property, i.e., having angle Θ\Theta close to π2\frac{\pi}{2} or 3π2\frac{3\pi}{2}, which is not far away from the mid-point between source and destination, i.e., 𝑿𝑿mid\left\|\boldsymbol{X}\right\|\approx\left\|\boldsymbol{X}_{\rm mid}\right\|, as the relay intensity increases.

For 𝑿Φ\boldsymbol{X}\in\Phi having the norm Ψ\Psi and angle Θ\Theta, we can express the relay selection function s^(𝑿)\widehat{s}\left(\boldsymbol{X}\right) as s^(𝑿)=Ψ2+2dΨ|cos(Θ)|+d2\widehat{s}\left(\boldsymbol{X}\right)=\sqrt{\Psi^{2}+2d\Psi\left|\cos\left(\Theta\right)\right|+d^{2}}. Hence, as the separation between source and destination nodes increases, the relay locations having angles close to 0 or π\pi are penalized more strongly. Since 𝒫mid\mathcal{P}_{\rm mid} chooses the relay node based only on the distance to mid-point criterion without considering the distance balancing dimension, it becomes less probable for a relay node selected by 𝒫mid\mathcal{P}_{\rm mid} to be optimum when dd is large. Correspondingly, the sufficient condition for the optimality of the mid-point relay selection policy is most useful when both the relay intensity and separation between source and destination nodes are small. Overall, despite its analytical convenience for performance evaluation, we observe that 𝒫mid\mathcal{P}_{\rm mid} suffers from its “distance balancing ignorant operation” notably, which provides a motivation for an in-depth search for the statistical properties of 𝒫opt\mathcal{P}_{\rm opt} in the remainder of the paper.

Remark 1: The performance gap between 𝒫opt\mathcal{P}_{\rm opt} and 𝒫mid\mathcal{P}_{\rm mid} can be significant in terms of average rate and outage probability depending on the network configuration and signal propagation characteristics. This point will be illustrated numerically in Section VIII.

Remark 2: Using the results in this section and 𝖱ave(𝒫mid){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm mid}\right), we can readily obtain an upper bound on 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right). This upper bound is given by

𝖱ave(𝒫opt)𝖱ave(𝒫mid)+Δave,\displaystyle{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right)\leq{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm mid}\right)+\Delta_{\rm ave}, (19)

where Δave\Delta_{\rm ave} is equal to

Δave=2𝖤[(1eλP(𝑿mid))log2(1+𝖲𝖭𝖱G(Ψmid2+d2)1+𝖲𝖭𝖱G(s^(𝑿mid)))𝟣{0Θmidπ2}]\displaystyle\Delta_{\rm ave}=2\mathsf{E}\left[\left(1-{\rm e}^{-\lambda P\left(\boldsymbol{X}_{\rm mid}\right)}\right)\log_{2}\left(\frac{1+{\sf SNR}\cdot G\left(\sqrt{\Psi_{\rm mid}^{2}+d^{2}}\right)}{1+{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\right)}\right)\mathsf{1}_{\left\{0\leq\Theta_{\rm mid}\leq\frac{\pi}{2}\right\}}\right]

for the no-fading scenario, and it is equal to

Δave=2ln2𝖤[(1eλP(𝑿mid))(f(1𝖲𝖭𝖱G(Ψmid2+d2))\displaystyle\Delta_{\rm ave}=\frac{2}{\ln 2}\mathsf{E}\left[\left(1-{\rm e}^{-\lambda P\left(\boldsymbol{X}_{\rm mid}\right)}\right)\left(f\left(\frac{1}{{\sf SNR}\cdot G\left(\sqrt{\Psi_{\rm mid}^{2}+d^{2}}\right)}\right)\right.\right.
f(1𝖲𝖭𝖱G(s^(𝑿mid))))𝟣{0Θmidπ2}]\displaystyle\left.\left.-f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\right)}\right)\rule{0.0pt}{25.6073pt}\right)\mathsf{1}_{\left\{0\leq\Theta_{\rm mid}\leq\frac{\pi}{2}\right\}}\right]

for the Rayleigh fading scenario with the function P(𝒙)P\left(\boldsymbol{x}\right) given as in Theorem 3 and the function f(x)f(x) defined as f(x)exE1(x)f(x)\triangleq{\rm e}^{x}\text{E}_{1}(x), where E1(x)\text{E}_{1}(x) is the exponential integral given by the identity E1(x)=1etxtdt\text{E}_{1}(x)=\int_{1}^{\infty}\frac{{\rm e}^{-tx}}{t}{\rm d}t for x>0x>0. For the Rayleigh fading case, we take HH as a circularly symmetric complex Gaussian random variable with unit power, i.e., H𝒞𝒩(0,1)H\sim\mathcal{CN}(0,1). Hence, |H|2\left|H\right|^{2} is an exponential random variable with unit mean, i.e., f|H|2(x)=exf_{\left|H\right|^{2}}(x)={\rm e}^{-x}. The derivations for the upper bounds above are given in Appendix D. We note that we can calculate 𝖱ave(𝒫mid){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm mid}\right) by averaging (16) with respect to the joint magnitude and angle distribution of 𝑿mid\boldsymbol{X}_{\rm mid}. Similarly, we can calculate Δave\Delta_{\rm ave} by taking above expectations with respect to the joint magnitude and angle distribution of 𝑿mid\boldsymbol{X}_{\rm mid}. In the next section, we will characterize the probability distribution of s^(𝑿opt)\widehat{s}\left(\boldsymbol{X}_{\rm opt}\right) and obtain an exact expression for 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right).

VI Performance Analysis for the Optimum Relay Selection Policy

In this section, we will obtain analytical expressions for 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) and 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right). To this end, we will take random relay location process Φ\Phi as an HPPP having intensity λ>0\lambda>0. Due to the stationarity and isotropy of HPPPs [76], we will assume that 𝒙s=(d,0)\boldsymbol{x}_{\rm s}=\left(-d,0\right)^{\top} and 𝒙d=(d,0)\boldsymbol{x}_{\rm d}=\left(d,0\right)^{\top} without loss of generality. This is the example network configuration illustrated in Fig. 1.

VI-A Distribution of s^(𝐗opt)\widehat{s}\left(\boldsymbol{X}_{\rm opt}\right)

As explained after the proof of Lemma 1, the conditional data rate, given the relay locations, is equal to

𝖱Φ(𝒫)=12𝖤[log2(1+𝖲𝖭𝖱|H|2G(s^(𝑿𝒫)))|Φ]\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}\right)=\frac{1}{2}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H\right|^{2}G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)\right)\big{|}\Phi\right] (20)

for any relay selection policy. Hence, the key step to obtain analytical expressions for 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) and 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) is to derive the distribution function for s^(𝑿opt)\widehat{s}\left(\boldsymbol{X}_{{\rm opt}}\right). For notational simplicity, we define Γopts^(𝑿opt)\Gamma_{\rm opt}\triangleq\widehat{s}\left(\boldsymbol{X}_{{\rm opt}}\right). Then, we have

Γopt=min𝑿Φs^(𝑿)\displaystyle\Gamma_{\rm opt}=\min_{\boldsymbol{X}\in\Phi}\widehat{s}\left(\boldsymbol{X}\right) (21)

by definition of 𝒫opt\mathcal{P}_{\rm opt}. That is, Γopt\Gamma_{\rm opt} is the minimum value achieved by the relay selection function over Φ\Phi. In the next theorem, we provide the cumulative distribution function (cdf) and the pdf of Γopt\Gamma_{\rm opt}.

Theorem 4

The cdf FΓopt(γ)F_{\Gamma_{\rm opt}}\left(\gamma\right) and the pdf fΓopt(γ)f_{\Gamma_{\rm opt}}\left(\gamma\right) of Γopt\Gamma_{\rm opt} are given by

FΓopt(γ)={0 if γ<d1e2λd2((γd)2arcsec(γd)(γd)21) if γd\begin{split}F_{\Gamma_{\rm opt}}\left(\gamma\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ 1-{\rm e}^{-2\lambda d^{2}\left(\left(\frac{\gamma}{d}\right)^{2}\operatorname{arcsec}\left(\frac{\gamma}{d}\right)-\sqrt{\left(\frac{\gamma}{d}\right)^{2}-1}\right)}&\mbox{ if }\gamma\geq d\end{array}\right.\end{split} (22)

and

fΓopt(γ)=4λγarcsec(γd)e2λd2((γd)2arcsec(γd)(γd)21)𝟣{γd}.\begin{split}f_{\Gamma_{\rm opt}}\left(\gamma\right)&=4\lambda\gamma\operatorname{arcsec}\left(\frac{\gamma}{d}\right){\rm e}^{-2\lambda d^{2}\left(\left(\frac{\gamma}{d}\right)^{2}\operatorname{arcsec}\left(\frac{\gamma}{d}\right)-\sqrt{\left(\frac{\gamma}{d}\right)^{2}-1}\right)}\mathsf{1}_{\left\{\gamma\geq d\right\}}\end{split}. (23)
Proof:

See Appendix E. ∎

There are several important remarks worth to mention about the functional forms of FΓoptF_{\Gamma_{\rm opt}} and fΓoptf_{\Gamma_{\rm opt}} given in Theorem 4. Let g(x)g(x) be the function defined as g(x)=x2arcsec(x)x21g(x)=x^{2}\operatorname{arcsec}(x)-\sqrt{x^{2}-1} for x1x\geq 1. It is easy to see that g(1)=0g(1)=0 and g(x)=2xarcsec(x)g^{\prime}(x)=2x\operatorname{arcsec}(x), which is the first derivative of g(x)g(x) with respect to xx. Since g(x)>0g^{\prime}(x)>0 for all x>1x>1, we conclude that g(x)g(x) is a strictly increasing and positive function of xx over (1,)\left(1,\infty\right). Further, it can also be seen that limxg(x)x2=π2\lim_{x\rightarrow\infty}\frac{g(x)}{x^{2}}=\frac{\pi}{2}. This shows that the exponents in (22) and (23) are non-positive for all values of γd\gamma\geq d with the tail of fΓoptf_{\Gamma_{\rm opt}} decaying according to limγln(fΓopt(γ))γ2=πλ\lim_{\gamma\rightarrow\infty}-\frac{\ln\left(f_{\Gamma_{\rm opt}}\left(\gamma\right)\right)}{\gamma^{2}}=\pi\lambda. Hence, all the moments of Γopt\Gamma_{\rm opt} exist although they may not be calculated in closed form. For example, 𝖤[Γopt]\mathsf{E}\left[\Gamma_{\rm opt}\right] can be obtained numerically by using (22) as below

𝖤[Γopt]\displaystyle\mathsf{E}\left[\Gamma_{\rm opt}\right] =\displaystyle= 0𝖯𝗋{Γopt>γ}dγ\displaystyle\int_{0}^{\infty}\mathsf{Pr}\left\{\Gamma_{\rm opt}>\gamma\right\}{\rm d}\gamma
=\displaystyle= d1e2λd2(γ2arcsec(γ)γ21)dγ,\displaystyle d\int_{1}^{\infty}{\rm e}^{-2\lambda d^{2}\left(\gamma^{2}\operatorname{arcsec}\left(\gamma\right)-\sqrt{\gamma^{2}-1}\right)}{\rm d}\gamma,

which cannot be reduced to a closed form.

Secondly, since g(x)>0g(x)>0 for all x>1x>1, it also holds that limλFΓopt(γ)=1\lim_{\lambda\rightarrow\infty}F_{\Gamma_{\rm opt}}\left(\gamma\right)=1 for γ>d\gamma>d and limλFΓopt(γ)=0\lim_{\lambda\rightarrow\infty}F_{\Gamma_{\rm opt}}\left(\gamma\right)=0 for γd\gamma\leq d. Therefore, Γopt\Gamma_{\rm opt} converges in distribution to a deterministic variable having value dd. This behaviour coincides with our intuition that there is always a relay in any disc with positive radius centered around the origin when the relay intensity becomes large. Indeed, it can be shown that Γopt\Gamma_{\rm opt} converges to dd almost surely as λ\lambda grows without a bound by using Borel-Cantelli lemma [77].

Finally, the proof given in Appendix E for Theorem 4 also reveals the pre-limit distribution for the minimum of relay selection function in the finite network case. Let Γopt,τ=min𝑿Φ(𝟎,τ)s^(𝑿)\Gamma_{{\rm opt},\tau}=\min_{\boldsymbol{X}\in\Phi\cap\mathcal{B}\left(\boldsymbol{0},\tau\right)}\widehat{s}\left(\boldsymbol{X}\right), where (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right) is the closed disc centered around the origin 𝟎\boldsymbol{0} and having radius τ\tau. Then, the cdf FΓopt,τ(γ)F_{\Gamma_{{\rm opt},\tau}}\left(\gamma\right) can be expressed as

FΓopt,τ(γ)=1eλπτ2FΓ(γ),\displaystyle F_{\Gamma_{{\rm opt},\tau}}\left(\gamma\right)=1-{\rm e}^{-\lambda\pi\tau^{2}F_{\Gamma}\left(\gamma\right)},

where FΓ(γ)F_{\Gamma}(\gamma) is given by (110) in Appendix D. The convergence of FΓopt,τ(γ)F_{\Gamma_{{\rm opt},\tau}}\left(\gamma\right) to FΓopt(γ)F_{\Gamma_{\rm opt}}\left(\gamma\right) as τ\tau increases is shown in Fig. 3, which is quite quick.

The optimum multi-user multi-relay selection problem for random spatial networks with interference is not tractable due to coupling between local relay selection processes at the active interfering links. However, the distribution convergence results shown above hints at an efficient sub-optimum heuristic for multi-user multi-relay selection in the presence of interference. If the layer 2 MAC protocol can achieve a certain minimum separation, sufficiently larger than the typical inter-relay distances, between interfering active transmitter-receiver pairs, then the effect of interference can be mitigated to a large extent in the relay selection process and the local relay selection processes can be decoupled from each other. That is, the differential interference at the optimum relay node (chosen by ignoring interference) and at other relays stays below a certain threshold with high probability to overcome the distance advantage provided by the interference-free optimum relay position when the MAC protocol can sufficiently mitigate interference in the vicinity of active communication links. With the suggested decoupling, the results in our paper can be directly applied to approximate the rate and outage performance of the optimum multi-user multi-relay selection policy by running locally optimum relay selection policies, restricted to a small region around each active link as in the interference-free case. This is specifically a co-design issue of MAC and relay selection algorithms, whose complete analysis is not within the scope of this paper.

Refer to captionRefer to caption
Figure 3: Convergence to the limiting case.

As seen in (20), an important determinant of the system performance is the random variable Sopt=𝖲𝖭𝖱G(Γopt)S_{\rm opt}={\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right). For a non-negative non-increasing path-loss function, the cdf of SoptS_{\rm opt} can be written as

FSopt(s)=1FΓopt(G1(s𝖲𝖭𝖱)),\displaystyle F_{S_{\rm opt}}(s)=1-F_{\Gamma_{\rm opt}}\left(G^{-1}\left(\frac{s}{{\sf SNR}}\right)\right), (24)

where G1(s)G^{-1}(s) is defined as G1(s)=inf{x0:G(x)s}G^{-1}(s)=\inf\left\{x\geq 0:G(x)\leq s\right\}. We note that FSopt(s)=1F_{S_{\rm opt}}(s)=1 for s𝖲𝖭𝖱G(d)s\geq{\sf SNR}\cdot G(d) since GG is non-increasing and Γoptd\Gamma_{\rm opt}\geq d. We will make use of FSopt(s)F_{S_{\rm opt}}(s) to calculate 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) below. Further, if GG is monotone decreasing and continuous, the pdf of SoptS_{\rm opt} can be written as

fSopt(s)=1𝖲𝖭𝖱fΓopt(G1(s𝖲𝖭𝖱))|1G(G1(s𝖲𝖭𝖱))|,\displaystyle f_{S_{\rm opt}}(s)=\frac{1}{{\sf SNR}}f_{\Gamma_{\rm opt}}\left(G^{-1}\left(\frac{s}{{\sf SNR}}\right)\right)\left|\frac{1}{G^{\prime}\left(G^{-1}\left(\frac{s}{{\sf SNR}}\right)\right)}\right|, (25)

which can used to calculate 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) for the power-law decaying path-loss models such as G(x)=1xαG(x)=\frac{1}{x^{\alpha}} or G(x)=11+xαG(x)=\frac{1}{1+x^{\alpha}} [78, 79, 80, 81, 82, 83].

VI-B Average Rate

Now, we provide expressions for the average rate achieved by the optimum relay selection policy 𝒫opt\mathcal{P}_{\rm opt}. We will consider both no-fading and Rayleigh fading cases. We recall that we consider a deterministic normalized fading gain with unit power for the no-fading case, and we take H𝒞𝒩(0,1)H\sim\mathcal{CN}(0,1) for the the Rayleigh fading case.

For 𝖱Φ(𝒫opt){\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right), we can write

𝖱Φ(𝒫opt)\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right) =\displaystyle= 12𝖤[log2(1+𝖲𝖭𝖱|H|2G(Γopt))|Φ]\displaystyle\frac{1}{2}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H\right|^{2}G\left(\Gamma_{\rm opt}\right)\right)\Big{|}\Phi\right] (28)
=\displaystyle= {12log2(1+𝖲𝖭𝖱G(Γopt))if no-fading12ln2e1𝖲𝖭𝖱G(Γopt)E1(1𝖲𝖭𝖱G(Γopt))if Rayleigh fading,\displaystyle\left\{\begin{array}[]{ll}\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)\right)&\text{if no-fading}\\ \frac{1}{2\ln 2}{\rm e}^{\frac{1}{{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)}}\text{E}_{1}\left(\frac{1}{{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)}\right)&\text{if Rayleigh fading}\end{array},\right.

where E1(x)=1etxtdt\text{E}_{1}(x)=\int_{1}^{\infty}\frac{{\rm e}^{-tx}}{t}{\rm d}t, x>0x>0, is the exponential integral as defined in Section V. The expression for 𝖱Φ(𝒫){\sf R}_{\Phi}\left(\mathcal{P}\right) for any relay selection policy 𝒫\mathcal{P} in the Rayleigh fading case was obtained in Appendix D while deriving an upper bound on 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right), which follows from integration-by-parts and change of variables. From Definition 3, the average rate, averaged over relay locations, can be calculated as

𝖱ave(𝒫opt)={12dlog2(1+𝖲𝖭𝖱G(γ))fΓopt(γ)dγif no-fading12ln2de1𝖲𝖭𝖱G(γ)E1(1𝖲𝖭𝖱G(γ))fΓopt(γ)dγif Rayleigh fading,\begin{split}{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right)&=\left\{\begin{array}[]{ll}\frac{1}{2}\int_{d}^{\infty}\log_{2}\left(1+{\sf SNR}\cdot G(\gamma)\right)f_{\Gamma_{\rm opt}}(\gamma){\rm d}\gamma&\text{if no-fading}\\ \frac{1}{2\ln 2}\int_{d}^{\infty}{\rm e}^{\frac{1}{{\sf SNR}\cdot G(\gamma)}}\text{E}_{1}\left(\frac{1}{{\sf SNR}\cdot G(\gamma)}\right)f_{\Gamma_{\rm opt}}(\gamma){\rm d}\gamma&\text{if Rayleigh fading}\end{array},\right.\end{split} (29)

where fΓoptf_{\Gamma_{\rm opt}} is as given in (23). On the other hand, for a monotone decreasing and continuous path-loss function, 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) can be expressed according to

𝖱ave(𝒫opt)={120𝖲𝖭𝖱G(d)log2(1+s)fSopt(s)dsif no-fading12ln20𝖲𝖭𝖱G(d)e1sE1(1s)fSopt(s)dsif Rayleigh fading\begin{split}{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right)&=\left\{\begin{array}[]{ll}\frac{1}{2}\int_{0}^{{\sf SNR}\cdot G\left(d\right)}\log_{2}\left(1+s\right)f_{S_{\rm opt}}(s){\rm d}s&\text{if no-fading}\\ \frac{1}{2\ln 2}\int_{0}^{{\sf SNR}\cdot G\left(d\right)}{\rm e}^{\frac{1}{s}}\text{E}_{1}\left(\frac{1}{s}\right)f_{S_{\rm opt}}(s){\rm d}s&\text{if Rayleigh fading}\end{array}\right.\end{split} (30)

by using fSopt(s)f_{S_{\rm opt}}(s) given in (25). To the best of our knowledge, it is not possible to reduce the integral expressions in (29) and (30) for 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) to a closed-form. However, these single integrals can be evaluated very quickly and efficiently by using standard numerical integration techniques.

VI-C Outage Probability

Recall that the outage probability 𝖯out(𝒫){\sf P}_{\rm out}\left(\mathcal{P}\right) for a relay selection policy 𝒫\mathcal{P} is defined to be the probability that the rate falls below a certain predetermined threshold ρ\rho. From Definition 2, we write 𝖯out(𝒫opt)=𝖯𝗋{𝖱Φ(𝒫opt)ρ}{\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right)=\mathsf{Pr}\left\{{\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right)\leq\rho\right\} for the optimum relay selection policy 𝒫opt\mathcal{P}_{\rm opt}. For the no-fading case, by using (28), we have

𝖯out(𝒫opt)\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) =\displaystyle= 𝖯𝗋{12log2(1+𝖲𝖭𝖱G(Γopt))ρ}\displaystyle\mathsf{Pr}\left\{\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)\right)\leq\rho\right\} (33)
=\displaystyle= {FSopt(22ρ1)if ρ<12log2(1+𝖲𝖭𝖱G(d))1if ρ12log2(1+𝖲𝖭𝖱G(d)),\displaystyle\left\{\begin{array}[]{ll}F_{S_{\rm opt}}\left(2^{2\rho}-1\right)&\mbox{if }\rho<\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G(d)\right)\\ 1&\text{if }\rho\geq\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G(d)\right)\end{array},\right.

where FSoptF_{S_{\rm opt}} is as given in (24). Similarly, for the Rayleigh fading case, we have

𝖯out(𝒫opt)\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) =\displaystyle= 𝖯𝗋{12ln2e1𝖲𝖭𝖱G(Γopt)E1(1𝖲𝖭𝖱G(Γopt))ρ}\displaystyle\mathsf{Pr}\left\{\frac{1}{2\ln 2}{\rm e}^{\frac{1}{{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)}}\text{E}_{1}\left(\frac{1}{{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)}\right)\leq\rho\right\} (36)
=\displaystyle= {FSopt(s)if s<𝖲𝖭𝖱G(d)1if s𝖲𝖭𝖱G(d),\displaystyle\left\{\begin{array}[]{ll}F_{S_{\rm opt}}\left(s^{\star}\right)&\mbox{if }s^{\star}<{\sf SNR}\cdot G\left(d\right)\\ 1&\text{if }s^{\star}\geq{\sf SNR}\cdot G\left(d\right)\end{array},\right.

where ss^{\star} is the unique solution for the equation 12ln2e1/sE1(1/s)=ρ\frac{1}{2\ln 2}{\rm e}^{1/s}\text{E}_{1}\left(1/s\right)=\rho. Uniqueness of ss^{\star} follows from the fact that the function f(x)=exE1(x)f(x)={\rm e}^{x}\text{E}_{1}\left(x\right) defined for x>0x>0 is a continuous and strictly decreasing function of xx, which attains the values limx0f(x)=\lim_{x\rightarrow 0}f(x)=\infty (i.e., this can be shown using monotone convergence theorem) and limxf(x)=0\lim_{x\rightarrow\infty}f(x)=0 (i.e., this can be shown using dominated convergence theorem). Hence, ss^{\star} can be readily calculated by using numerical methods such as the bisection technique or by using Matlab’s vpasolve routine. Moreover, by using the inequality 12ln(1+2x)<exE1(x)<ln(1+1x)\frac{1}{2}\ln\left(1+\frac{2}{x}\right)<{\rm e}^{x}\text{E}_{1}(x)<\ln\left(1+\frac{1}{x}\right) (i.e., see [84, p. 229, 5.1.20]), we can also bound 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) for the Rayleigh fading case according to FSopt(22ρ1)𝖯out(𝒫opt)FSopt(24ρ12)F_{S_{\rm opt}}\left(2^{2\rho}-1\right)\leq{\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right)\leq F_{S_{\rm opt}}\left(\frac{2^{4\rho}-1}{2}\right). The lower bound on 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) coincides with the one that can be obtained by using 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) for the no-fading case in (33).

VII Relay Selection with Selective Distributed-Feedback

In the previous sections, we have characterized key properties for the optimum relay selection policy, its outage and average rate performance as well as the optimality probability for the mid-point relay selection policy when the relay nodes are randomly distributed over the plane according to an HPPP of any given intensity. This analysis, however, assumes a centralized operation in which information pertaining to all relay locations is available at the source node (or at a central entity) to solve the optimum relay selection problem in (12). Even though the class of relay selection policies we investigate in this paper only requires relay locations as the CSI, which change at a much slower time-scale than fading, the task of feeding this information back to the source node is still onerous, and hence an impeding factor for physical deployments. One potential means of reducing the feedback load is to adopt a selective distributed-feedback relay selection policy in which relay nodes, whose locations are given according to Φ\Phi, are provided with the autonomy of giving their own feedback decisions independently from other relays in the network based on their local channel quality indicators s^(𝑿)\widehat{s}\left(\boldsymbol{X}\right) for 𝑿Φ\boldsymbol{X}\in\Phi.

VII-A Threshold Feedback Policies for Relay Selection and Their Statistical Properties

We will study simple but practical threshold feedback policies to regulate the feedback load in the network. This class of feedback polices possesses certain optimality properties to maximize data rates [85, 86, 87, 88, 89, 90, 91]. Here, we will utilize them to control the number of relay nodes feeding their channel states back to the source node as a measure of the total feedback load in the network. More explicitly, for any given threshold value T0T\geq 0, we will say that a relay node located at 𝑿Φ\boldsymbol{X}\in\Phi and operating according to a threshold-based selective distributed-feedback relay selection policy with threshold value TT will feed its channel quality indicator s^(𝑿)\widehat{s}\left(\boldsymbol{X}\right) back to the source node if and only if s^(𝑿)T\widehat{s}\left(\boldsymbol{X}\right)\leq T. Hence, the total number of relays feeding back is given by

NFB=𝑿Φ𝟣{s^(𝑿)T}.\displaystyle N_{\rm FB}=\sum_{\boldsymbol{X}\in\Phi}\mathsf{1}_{\left\{\widehat{s}\left(\boldsymbol{X}\right)\leq T\right\}}. (37)

The average number of relays feeding back is then equal to

μ(T)=𝖤[NFB].\displaystyle\mu\left(T\right)=\mathsf{E}\left[N_{\rm FB}\right]. (38)

The next theorem characterizes the distribution of NFBN_{\rm FB} and the functional form of its average value.

Theorem 5

For any given threshold value T0T\geq 0, NFBN_{\rm FB} is a Poisson distributed random variable whose mean μ(T)\mu\left(T\right) is given according to

μ(T)={0 if T<dλπT22dλT2d22T2λarctan(dT2d2) if Td.\displaystyle\mu\left(T\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }T<d\\ \lambda\pi T^{2}-2d\lambda\sqrt{T^{2}-d^{2}}-2T^{2}\lambda\arctan\left(\frac{d}{\sqrt{T^{2}-d^{2}}}\right)&\mbox{ if }T\geq d\end{array}\right.. (41)
Proof:

See Appendix F. ∎

Using (41), it can easily be seen that μ(d)=0\mu\left(d\right)=0 and μ(T)\mu\left(T\right) is a continuous function of TT. Further, by using (37) and (38), it can also be seen that μ(T)\mu\left(T\right) is a monotone increasing function of TT with limit limTμ(T)=\lim_{T\rightarrow\infty}\mu\left(T\right)=\infty. Hence, for any given feedback load μ00\mu_{0}\geq 0, we are guaranteed to find a threshold value T0T_{0} such that μ(T0)=μ0\mu\left(T_{0}\right)=\mu_{0} by intermediate value theorem. From a network design perspective, this observation shows that the class of threshold-based distributed relay selection policies is rich enough to satisfy any given feedback load constraint on the network by properly allocating a common threshold value to all relay nodes and allowing them to operate autonomously while giving their feedback decisions based on this common threshold value.

Refer to captionRefer to caption
Figure 4: Probability distribution of the number of relays feeding back for T=3T=3 and d=1d=1. (λ=0.5\lambda=0.5 for the left-hand side figure and λ=1\lambda=1 for the right-hand side figure.)
Refer to captionRefer to caption
Figure 5: Average number of relays feeding back and the probability of having at least one relay feeding back for d=1d=1 and various values of λ\lambda.

In Fig. 4, we plot the simulated distributions of NFBN_{\rm FB} for λ=0.5\lambda=0.5 and λ=1\lambda=1, and compare them with the Poisson distribution having mean μ(T)\mu\left(T\right). As predicted by Theorem 5, simulated and theoretical distributions match each other perfectly. In Fig. 5, we plot the average number of relays feeding back μ(T)\mu\left(T\right) and the probability of at least one relay feeding back 𝖯𝗋{NFB1}\mathsf{Pr}\left\{N_{\rm FB}\geq 1\right\} as a function of TT. Again, there is a perfect match between simulated and analytical curves, verifying the predictions in Theorem 5. In particular, 𝖯𝗋{NFB1}\mathsf{Pr}\left\{N_{\rm FB}\geq 1\right\} is an important performance indicator for relay-assisted spatial wireless networks with distributed relay selection. The relay with optimum location is always among the relays feeding their channel quality indicators back to the source node if NFB1N_{\rm FB}\geq 1. Hence, with probability 𝖯𝗋{NFB1}\mathsf{Pr}\left\{N_{\rm FB}\geq 1\right\}, there is no loss of optimality arising from implementing a threshold-based selective feedback distributed relay selection mechanism. Based on Theorem 5, this probability is equal to 𝖯𝗋{NFB1}=1eμ(T)\mathsf{Pr}\left\{N_{\rm FB}\geq 1\right\}=1-{\rm e}^{-\mu\left(T\right)}. This shows that the performance loss due to having a threshold-based selective feedback distributed relay selection mechanism diminishes exponentially fast as a function of the feedback load in the network, which we measure in terms of the average number of relays feeding their channel quality indicators back to the source node. Numerically, we have 𝖯𝗋{NFB1}0.99\mathsf{Pr}\left\{N_{\rm FB}\geq 1\right\}\leq 0.99 whenever μ(T)5\mu\left(T\right)\geq 5. As a result, for a given relay intensity and source-destination separation, choosing the threshold value such that μ(T)=5\mu\left(T\right)=5 implies almost a negligible performance loss, whilst providing a massive reduction in the total feedback load required to achieve 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) and 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right).

VII-B Average Rate and Outage Probability Achieved

Next, we obtain analytical expressions for the average rate and outage probability achieved by a threshold-based selective feedback distributed relay selection policy 𝒫FB\mathcal{P}_{\rm FB}, analogous to those obtained in Section VI. Different from previous parts, we will denote the average rate and outage probability by 𝖱ave(𝒫FB,T){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm FB},T\right) and 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) with a slight abuse of notation, respectively, to indicate their dependency on the threshold level TT. Due to the distributed relay selection mechanism, collection of relay nodes whose channel quality indicators are available at the source node is a random set with size NFBN_{\rm FB}. If NFB1N_{\rm FB}\geq 1, there is one or more relay nodes feeding back to the source node, and the source selects the best relay among them for data transmission. On the other hand, if NFB=0N_{\rm FB}=0, there is no relay feeding back to the source, and we assume that the source node does not transmit any data due to lack of CSI in this case. Since NFB=0N_{\rm FB}=0 for T<dT<d (i.e., no relay feeds back for this range of TT), we will only analyze the case TdT\geq d below. Under these operating conditions, the following theorem provides the analytical expressions for the average rate achieved by 𝒫FB\mathcal{P}_{\rm FB}.

Theorem 6

For a given threshold-based selective feedback distributed relay selection policy 𝒫FB\mathcal{P}_{\rm FB} having a threshold value TdT\geq d, the average rate 𝖱ave(𝒫FB,T){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm FB},T\right) is equal to

𝖱ave(𝒫FB,T)={12dTlog2(1+𝖲𝖭𝖱G(γ))fΓopt(γ)dγif no-fading12ln2dTe1𝖲𝖭𝖱G(γ)E1(1𝖲𝖭𝖱G(γ))fΓopt(γ)dγif Rayleigh fading,\displaystyle{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm FB},T\right)=\left\{\begin{array}[]{ll}\frac{1}{2}\int_{d}^{T}\log_{2}\left(1+{\sf SNR}\cdot G\left(\gamma\right)\right)f_{\Gamma_{\rm opt}}\left(\gamma\right){\rm d}\gamma&\mbox{if no-fading}\\ \frac{1}{2\ln 2}\int_{d}^{T}{\rm e}^{\frac{1}{{\sf SNR}\cdot G\left(\gamma\right)}}\text{E}_{1}\left(\frac{1}{{\sf SNR}\cdot G\left(\gamma\right)}\right)f_{\Gamma_{\rm opt}}\left(\gamma\right){\rm d}\gamma&\mbox{if Rayleigh fading}\end{array}\right., (44)

where fΓopt(γ)f_{\Gamma_{\rm opt}}(\gamma) is given as in Theorem 4.

Proof:

The proof follows from the equivalence of events {NFB1}\left\{N_{\rm FB}\geq 1\right\} and {ΓoptT}\left\{\Gamma_{\rm opt}\leq T\right\}, and writing the rate achieved by 𝒫FB\mathcal{P}_{\rm FB}, conditioned on the relay locations, according to

𝖱Φ(𝒫FB,T)=12log2(1+𝖲𝖭𝖱G(Γopt))𝟣{ΓoptT}\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}_{\rm FB},T\right)=\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)\right)\mathsf{1}_{\left\{\Gamma_{\rm opt}\leq T\right\}}

when there is no fading, and according to

𝖱Φ(𝒫FB,T)=12ln2e1𝖲𝖭𝖱G(Γopt)E1(1𝖲𝖭𝖱G(Γopt))𝟣{ΓoptT}\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}_{\rm FB},T\right)=\frac{1}{2\ln 2}{\rm e}^{\frac{1}{{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)}}\text{E}_{1}\left(\frac{1}{{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)}\right)\mathsf{1}_{\left\{\Gamma_{\rm opt}\leq T\right\}}

when there is Rayleigh fading by using (28). ∎

We note that (44) can be written as

𝖱ave(𝒫FB,T)={12𝖲𝖭𝖱G(T)𝖲𝖭𝖱G(d)log2(1+s)fSopt(s)dsif no-fading12ln2𝖲𝖭𝖱G(T)𝖲𝖭𝖱G(d)e1sE1(1s)fSopt(s)dsif Rayleigh fading,\displaystyle{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm FB},T\right)=\left\{\begin{array}[]{ll}\frac{1}{2}\int_{{\sf SNR}\cdot G\left(T\right)}^{{\sf SNR}\cdot G(d)}\log_{2}\left(1+s\right)f_{S_{\rm opt}}\left(s\right){\rm d}s&\mbox{if no-fading}\\ \frac{1}{2\ln 2}\int_{{\sf SNR}\cdot G\left(T\right)}^{{\sf SNR}\cdot G(d)}{\rm e}^{\frac{1}{s}}\text{E}_{1}\left(\frac{1}{s}\right)f_{S_{\rm opt}}\left(s\right){\rm d}s&\mbox{if Rayleigh fading}\end{array}\right., (47)

where fSopt(s)f_{S_{\rm opt}}(s) is given as in (25) for monotone decreasing and continuous path-loss functions. In the next theorem, we provide similar expressions for the outage probability 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) achieved by 𝒫FB\mathcal{P}_{\rm FB}.

Theorem 7

For a given threshold-based selective feedback distributed relay selection policy 𝒫FB\mathcal{P}_{\rm FB} having a threshold value TdT\geq d, the outage probability 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) is equal to

𝖯out(𝒫FB,T)\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)
={eμ(T) if ρ12log2(1+𝖲𝖭𝖱G(T))FSopt(22ρ1) if 12log2(1+𝖲𝖭𝖱G(T))<ρ<12log2(1+𝖲𝖭𝖱G(d))1 if ρ12log2(1+𝖲𝖭𝖱G(d))\displaystyle=\left\{\begin{array}[]{ll}{\rm e}^{-\mu\left(T\right)}&\mbox{ if }\rho\leq\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(T\right)\right)\\ F_{S_{\rm opt}}\left(2^{2\rho}-1\right)&\mbox{ if }\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(T\right)\right)<\rho<\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(d\right)\right)\\ 1&\mbox{ if }\rho\geq\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(d\right)\right)\end{array}\right. (51)

when there is no fading, where FSoptF_{S_{\rm opt}} is given as in (24). On the other hand, for Rayleigh distributed fading, 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) is equal to

𝖯out(𝒫FB,T)={eμ(T) if s𝖲𝖭𝖱G(T)FSopt(s) if 𝖲𝖭𝖱G(T)<s<𝖲𝖭𝖱G(d)1 if s𝖲𝖭𝖱G(d),\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)=\left\{\begin{array}[]{ll}{\rm e}^{-\mu\left(T\right)}&\mbox{ if }s^{\star}\leq{\sf SNR}\cdot G\left(T\right)\\ F_{S_{\rm opt}}\left(s^{\star}\right)&\mbox{ if }{\sf SNR}\cdot G\left(T\right)<s^{\star}<{\sf SNR}\cdot G\left(d\right)\\ 1&\mbox{ if }s^{\star}\geq{\sf SNR}\cdot G\left(d\right)\end{array},\right. (55)

where ss^{\star} is the unique solution of the equation 12ln2e1/sE1(1/s)=ρ\frac{1}{2\ln 2}{\rm e}^{1/s}\text{E}_{1}\left(1/s\right)=\rho.

Proof:

We first consider the no-fading case. In this case, we have

𝖯out(𝒫FB,T)\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) =\displaystyle= 𝖯𝗋{𝖱Φ(𝒫FB,T)ρ}\displaystyle\mathsf{Pr}\left\{{\sf R}_{\Phi}\left(\mathcal{P}_{\rm FB},T\right)\leq\rho\right\} (56)
=\displaystyle= 𝖯𝗋{12log2(1+𝖲𝖭𝖱G(Γopt))𝟣{ΓoptT}ρ}.\displaystyle\mathsf{Pr}\left\{\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)\right)\mathsf{1}_{\left\{\Gamma_{\rm opt}\leq T\right\}}\leq\rho\right\}.

Using (56), it can be seen that we can write the outage event as the union of two disjoint events according to

{𝖱Φ(𝒫FB,T)ρ}={Γopt>T}({ΓoptT}{12log2(1+𝖲𝖭𝖱G(Γopt))ρ}).\displaystyle\left\{{\sf R}_{\Phi}\left(\mathcal{P}_{\rm FB},T\right)\leq\rho\right\}=\left\{\Gamma_{\rm opt}>T\right\}\bigcup\left(\left\{\Gamma_{\rm opt}\leq T\right\}\bigcap\left\{\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)\right)\leq\rho\right\}\right).

Hence, 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) is equal to

𝖯out(𝒫FB,T)\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) =\displaystyle= 𝖯𝗋{Γopt>T}+𝖯𝗋({ΓoptT}{12log2(1+𝖲𝖭𝖱G(Γopt))ρ})\displaystyle\mathsf{Pr}\left\{\Gamma_{\rm opt}>T\right\}+\mathsf{Pr}\left(\left\{\Gamma_{\rm opt}\leq T\right\}\bigcap\left\{\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)\right)\leq\rho\right\}\right) (57)
=\displaystyle= eμ(T)+𝖯𝗋({ΓoptT}{12log2(1+𝖲𝖭𝖱G(Γopt))ρ})\displaystyle{\rm e}^{-\mu\left(T\right)}+\mathsf{Pr}\left(\left\{\Gamma_{\rm opt}\leq T\right\}\bigcap\left\{\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)\right)\leq\rho\right\}\right)
=\displaystyle= eμ(T)+𝖯𝗋{G1(22ρ1𝖲𝖭𝖱)ΓoptT}.\displaystyle{\rm e}^{-\mu\left(T\right)}+\mathsf{Pr}\left\{G^{-1}\left(\frac{2^{2\rho}-1}{{\sf SNR}}\right)\leq\Gamma_{\rm opt}\leq T\right\}.

If G1(22ρ1𝖲𝖭𝖱)TG^{-1}\left(\frac{2^{2\rho}-1}{{\sf SNR}}\right)\geq T, then the second term in (57) is equal to zero, and we have 𝖯out(𝒫FB,T)=eμ(T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)={\rm e}^{-\mu\left(T\right)}. This case corresponds to ρ12log2(1+𝖲𝖭𝖱G(T))\rho\leq\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(T\right)\right), which is the first condition in (51). If d<G1(22ρ1𝖲𝖭𝖱)<Td<G^{-1}\left(\frac{2^{2\rho}-1}{{\sf SNR}}\right)<T, we have 𝖯out(𝒫FB,T)=𝖯𝗋{ΓoptG1(22ρ1𝖲𝖭𝖱)}=FSopt(22ρ1){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)=\mathsf{Pr}\left\{\Gamma_{\rm opt}\geq G^{-1}\left(\frac{2^{2\rho}-1}{{\sf SNR}}\right)\right\}=F_{S_{\rm opt}}\left(2^{2\rho}-1\right). This case corresponds to the second condition in (51). Finally, if G1(22ρ1𝖲𝖭𝖱)dG^{-1}\left(\frac{2^{2\rho}-1}{{\sf SNR}}\right)\leq d, we have 𝖯out(𝒫FB,T)=𝖯𝗋{ΓoptG1(22ρ1𝖲𝖭𝖱)}=1{\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)=\mathsf{Pr}\left\{\Gamma_{\rm opt}\geq G^{-1}\left(\frac{2^{2\rho}-1}{{\sf SNR}}\right)\right\}=1 since Γopt\Gamma_{\rm opt} is always greater than dd. This case corresponds to ρ12log2(1+𝖲𝖭𝖱G(d))\rho\geq\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(d\right)\right), which is the the third condition in (51).

For the Rayleigh fading case, we have

𝖯out(𝒫FB,T)\displaystyle{\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) =\displaystyle= eμ(T)+𝖯𝗋({ΓoptT}{12ln2f(1𝖲𝖭𝖱G(Γopt))ρ}),\displaystyle{\rm e}^{-\mu\left(T\right)}+\mathsf{Pr}\left(\left\{\Gamma_{\rm opt}\leq T\right\}\bigcap\left\{\frac{1}{2\ln 2}f\left(\frac{1}{{\sf SNR}\cdot G\left(\Gamma_{\rm opt}\right)}\right)\leq\rho\right\}\right), (58)

where f(x)f(x) is as defined in Section V, i.e., f(x)exE1(x)f(x)\triangleq{\rm e}^{x}\text{E}_{1}(x) for x>0x>0. As explained earlier in the paper, f(x)f(x) is a continuous and strictly decreasing function of xx with limiting values limx0f(x)=\lim_{x\rightarrow 0}f(x)=\infty and limxf(x)=0\lim_{x\rightarrow\infty}f(x)=0. Hence, defining γ\gamma^{\star} as the value for the first time f(1𝖲𝖭𝖱G(γ))f\left(\frac{1}{{\sf SNR}\cdot G\left(\gamma\right)}\right) is above 2ρln22\rho\ln 2, i.e., γinf{γ>0:f(1𝖲𝖭𝖱G(γ))2ρln2}\gamma^{\star}\triangleq\inf\left\{\gamma>0:f\left(\frac{1}{{\sf SNR}\cdot G\left(\gamma\right)}\right)\geq 2\rho\ln 2\right\}, analyzing three cases γd\gamma^{\star}\leq d, d<γ<Td<\gamma^{\star}<T and γT\gamma^{\star}\geq T separately, and defining s𝖲𝖭𝖱G(γ)s^{\star}\triangleq{\sf SNR}\cdot G\left(\gamma^{\star}\right), we obtain (55). ∎

Several important remarks are in order about Theorem 7 characterizing the outage probability achievable with a threshold-based selective feedback distributed relay selection policy. In particular, we observe two regimes emerging in Theorem 7 when we vary the target rate for both with and without fading. When ρ12log2(1+𝖲𝖭𝖱G(T))\rho\leq\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(T\right)\right) without fading or s𝖲𝖭𝖱G(T)s^{\star}\leq{\sf SNR}\cdot G\left(T\right) with fading where ss^{\star} is the solution for 12ln2e1/sE1(1/s)=ρ\frac{1}{2\ln 2}{\rm e}^{1/s}\text{E}_{1}\left(1/s\right)=\rho, the outage probability is equal to 𝖯out(𝒫FB,T)=eμ(T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)={\rm e}^{-\mu\left(T\right)}. This is the feedback-limited regime in which 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) is the same for both no-fading and fading cases, depends only on the average number of relays feeding back (which in turn depends on TT, λ\lambda and dd), and is independent of the target rate and fading behaviour. In this regime, the threshold value is set so small that we are guaranteed to achieve the target rate whenever there is at least one relay feeding its channel quality indicator back to the source node. We recall that the smaller TT is, the better CQI relays feeding back have, and achieve higher rates.

The second regime is the rate-limited regime that emerges when 12log2(1+𝖲𝖭𝖱G(T))<ρ<12log2(1+𝖲𝖭𝖱G(d))\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(T\right)\right)<\rho<\frac{1}{2}\log_{2}\left(1+{\sf SNR}\cdot G\left(d\right)\right) for the no-fading case and when 𝖲𝖭𝖱G(T)<s<𝖲𝖭𝖱G(d){\sf SNR}\cdot G\left(T\right)<s^{\star}<{\sf SNR}\cdot G\left(d\right) for the fading case. In this regime, the outage probability is equal to 𝖯out(𝒫FB,T)=FSopt(22ρ1){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)=F_{S_{\rm opt}}\left(2^{2\rho}-1\right) for the no-fading case and 𝖯out(𝒫FB,T)=FSopt(s){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)=F_{S_{\rm opt}}\left(s^{\star}\right) for the fading case, and hence 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) is a function of ρ\rho, the same for the all-feedback and selective feedback cases, and independent of TT. 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) also depends on the fading behaviour since ss^{\star} is not necessarily the same with 22ρ12^{2\rho}-1. In this regime, the threshold value is set so big that we are guaranteed to have at least one relay feeding its CQI back to the source node whenever any relay achieves the target rate.

VIII Numerical Results and Discussion

In this section, we will present our numerical results in order to illustrate the performance of 𝒫opt\mathcal{P}_{\rm opt} with and without feedback limitations. All distances are normalized to a unit distance. The path-loss exponent α\alpha is set to 44, and the path-loss function is taken to be G(x)=1x4G(x)=\frac{1}{x^{4}}. For the fading scenario, we consider Rayleigh fading coefficients with unit power. To benchmark the performance of the optimum relay selection policy, we also consider the network performance when the relay is chosen in such a way that it is closest to the mid-point and to the destination.555The performance of the relay selection policy choosing the relay closest to the source is the same with the performance of the one choosing the relay closest to the destination due to symmetry in the problem. Hence, the performance curves pertaining to the former one are not included.

The performance curves for the mid-point and closest-to-destination policies are obtained by using the distribution functions given in (59)-(62), where Γmids^(𝑿mid)\Gamma_{\rm mid}\triangleq\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right), ΓC2Ds^(𝑿C2D)\Gamma_{\rm C2D}\triangleq\widehat{s}\left(\boldsymbol{X}_{\rm C2D}\right), 𝑿C2D\boldsymbol{X}_{\rm C2D} is the relay location closest to the destination, and FNN(ψ)=(1eλπψ2)𝟣{ψ0}F_{\rm NN}\left(\psi\right)=\left(1-{\rm e}^{-\lambda\pi\psi^{2}}\right)\mathsf{1}_{\left\{\psi\geq 0\right\}} is the nearest-neighbor cdf. The derivations are given in Appendix G.

FΓmid(γ)=(FNN(γ2d2)2λ(γd)2γ2d2eλπxarccos(γ2xd22dx)dx)𝟣{γd}\displaystyle F_{\Gamma_{\rm mid}}\left(\gamma\right)=\left(F_{\rm NN}\left(\sqrt{\gamma^{2}-d^{2}}\right)-2\lambda\int_{\left(\gamma-d\right)^{2}}^{\gamma^{2}-d^{2}}{\rm e}^{-\lambda\pi x}\arccos\left(\frac{\gamma^{2}-x-d^{2}}{2d\sqrt{x}}\right){\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}} (59)
fΓmid(γ)=(4λγ(γd)2γ2d2eλπx4d2x(γ2xd2)2dx)𝟣{γd}\displaystyle f_{\Gamma_{\rm mid}}\left(\gamma\right)=\left(4\lambda\gamma\int_{\left(\gamma-d\right)^{2}}^{\gamma^{2}-d^{2}}\frac{{\rm e}^{-\lambda\pi x}}{\sqrt{4d^{2}x-\left(\gamma^{2}-x-d^{2}\right)^{2}}}{\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}} (60)
FΓC2D(γ)=(FNN(γ2d)+λ(γ2d)2γ2eλπxarccos(x+4d2γ24dx)dx)𝟣{γd}\displaystyle F_{\Gamma_{\rm C2D}}\left(\gamma\right)=\left(F_{\rm NN}\left(\gamma-2d\right)+\lambda\int_{\left(\gamma-2d\right)^{2}}^{\gamma^{2}}{\rm e}^{-\lambda\pi x}\arccos\left(\frac{x+4d^{2}-\gamma^{2}}{4d\sqrt{x}}\right){\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}} (61)
fΓC2D(γ)=2λγ(arccos(dγ)eλπγ2+(γ2d)2γ2eλπx16d2x(x+4d2γ2)2dx)𝟣{γd}\displaystyle f_{\Gamma_{\rm C2D}}\left(\gamma\right)=2\lambda\gamma\left(\arccos\left(\frac{d}{\gamma}\right){\rm e}^{-\lambda\pi\gamma^{2}}+\int_{\left(\gamma-2d\right)^{2}}^{\gamma^{2}}\frac{{\rm e}^{-\lambda\pi x}}{\sqrt{16d^{2}x-\left(x+4d^{2}-\gamma^{2}\right)^{2}}}{\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}} (62)

We note that the change in scale in how we measure the distances will only scale the 𝖲𝖭𝖱{\sf SNR} values in our analysis below, without changing the emerging fundamental trends in the network performance. Some recent studies indicated that the power-law functions model the path-loss more accurately for distances larger than 350350 meters [92]. Hence, for the relay densities below, one can consider the unit distance to be on the order of kilometers for higher practical relevance of the presented results.

Refer to caption
(a) Outage probability vs λ\lambda.
Refer to caption
(b) Average rate vs λ\lambda.
Figure 6: Outage probability and average rate achieved by different relay selection strategies as a function of λ\lambda for 𝖲𝖭𝖱=5{\sf SNR}=5 dB, d=1d=1 and α=4\alpha=4.

Fig. 6 shows the outage probability and average rate curves as a function of λ\lambda for 𝖲𝖭𝖱=5{\sf SNR}=5 dB. We set ρ=0.5\rho=0.5 [bits/s/Hz] for the outage probability curves and d=1d=1. Fig. 6a shows that the closest-to-destination scheme has the worst and degrading performance, where the network is in outage with high probability, i.e., 𝖯out(𝒫)1{\sf P}_{\rm out}\left(\mathcal{P}\right)\rightarrow 1, when λ\lambda increases. The reason for this phenomenon is the myopic selection of the relay without attempting to balance the relay-to-source and relay-to-destination distances, a problem which is more exacerbated for dense relay deployments. This observation clearly shows the importance of the distance balancing property of the optimum relay selection policy as discussed in Section V, and how poorly a relay selection policy, which does not respect this property, can perform.

Secondly, we observe that the optimum relay selection can outperform the mid-point relay selection significantly in terms of outage performance. For example, at λ=3\lambda=3, the outage probability improvements of the optimum policy with respect to the mid-point one are around 4848% for the no-fading case and 3434% for the fading case. The slope of the outage probability in logarithmic scale shows how fast the outage probability decreases with respect to λ\lambda. In particular, logarithm of the outage probability decreases almost linearly with λ\lambda for both optimum and mid-point selection policies but with different slopes and a slight decrease in the slope with λ\lambda for the mid-point policy. For the optimum relay selection policy, the decay rate is around 0.350.35 for the no-fading case and 0.240.24 for the fading case. On the other hand, for the mid-point relay selection policy, those values are 0.240.24 for the no-fading case, and 0.140.14 for the fading case.

Figs. 6b shows the average rate as a function of λ\lambda. While we notice similar performance variations, we can see that there is a performance gap between optimum and mid-point relay selection. For example, we lose around 0.040.04 [bits/s/Hz] at λ=4\lambda=4 for both no-fading and Rayleigh fading cases by implementing mid-point relay selection. For optimum relay selection, the average rate monotonically increases with λ\lambda to the limiting values 12log2(1+𝖲𝖭𝖱dα)\frac{1}{2}\log_{2}\left(1+\frac{{\sf SNR}}{d^{\alpha}}\right) and 12ln2edα𝖲𝖭𝖱E1(dα𝖲𝖭𝖱)\frac{1}{2\ln 2}{\rm e}^{\frac{d^{\alpha}}{{\sf SNR}}}\text{E}_{1}\left(\frac{d^{\alpha}}{{\sf SNR}}\right) for the no-fading and Rayleigh fading cases, respectively. For closest-to-destination relay selection, the average rate first increases and then decreases with λ\lambda to the limiting values 12log2(1+𝖲𝖭𝖱2αdα)\frac{1}{2}\log_{2}\left(1+\frac{{\sf SNR}}{2^{\alpha}d^{\alpha}}\right) and 12ln2e2αdα𝖲𝖭𝖱E1(2αdα𝖲𝖭𝖱)\frac{1}{2\ln 2}{\rm e}^{\frac{2^{\alpha}d^{\alpha}}{{\sf SNR}}}\text{E}_{1}\left(\frac{2^{\alpha}d^{\alpha}}{{\sf SNR}}\right) for the no-fading and Rayleigh fading cases, respectively. The asymptotic values are at 1.031.03 [bits/s/Hz] for the no-fading case and 0.860.86 [bits/s/Hz] for the Rayleigh fading case with optimum relay selection; and they are at 0.130.13 [bits/s/Hz] for the no-fading case and 0.120.12 [bits/s/Hz] for the fading case with closest-to-destination relay selection.

Refer to caption
(a) Average rate vs λ\lambda without fading.
Refer to caption
(b) Average rate vs λ\lambda with fading.
Figure 7: Average rate achieved by the threshold-based selective feedback distributed relay selection policy for various values of the threshold level.

In Fig. 7, we plot the average rate 𝖱ave(𝒫FB,T){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm FB},T\right) achieved by the threshold-based selective feedback distributed relay selection policy 𝒫FB\mathcal{P}_{\rm FB} as a function of λ\lambda for various values of TdT\geq d, where we still set d=1d=1 and 𝖲𝖭𝖱=5 dB{\sf SNR}=5\mbox{ dB}. For small values of TT, there is a large gap between the average rates achieved by 𝒫FB\mathcal{P}_{\rm FB} and all-feedback policies. This is due to the fact that 𝖯𝗋{NFB=0}=eμ(T)\mathsf{Pr}\left\{N_{\rm FB}=0\right\}={\rm e}^{-\mu\left(T\right)} is large for small values of TT, and hence the source node cannot receive any CQI from relays to choose one with high probability. More specifically, for the considered range of λ[0.1,4]\lambda\in\left[0.1,4\right] in Fig. 7, μ(T)\mu\left(T\right) ranges from 0.01230.0123 to 0.49340.4934 for T=1.1T=1.1, which indicates that the source node is without any CQI more than 60%60\% of time even for the most crowded relay network scenario considered in this figure.

A similar behaviour with a decreased rate gap between the selective feedback and all-feedback cases continues to hold for T=1.25T=1.25. In this case, the source node cannot access to any CQI around 13%13\% of time for the most crowded relay network scenario. For T=1.5T=1.5, μ(T)\mu\left(T\right) is approximately equal to 3.13.1, 4.654.65 and 6.26.2 for λ=2,3\lambda=2,3 and 44, respectively. As observed in Fig. 7, the rate gap between the selective feedback and all-feedback cases becomes very small after λ2\lambda\geq 2 for T=1.5T=1.5, which implies a significant reduction in the feedback load without sacrificing from the achievable data rates. Similar observations but with a smaller rate gap continues to hold for T=2T=2.

Refer to caption
(a) Outage probability vs λ\lambda for ρ=0.3\rho=0.3.
Refer to caption
(b) Outage probability vs λ\lambda for ρ=0.3\rho=0.3.
Refer to caption
(c) Outage probability vs ρ\rho for λ=2\lambda=2.
Refer to caption
(d) Outage probability vs ρ\rho for λ=2\lambda=2.
Figure 8: Outage probability achieved by the threshold-based selective feedback distributed relay selection policy for various values of the threshold level.

In Fig. 8, we plot the outage probability curves achieved by 𝒫FB\mathcal{P}_{\rm FB} for both as a function of λ\lambda and ρ\rho. Feedback-limited and rate-limited regimes discussed after Theorem 7 for outage probability are also apparent in this figure. While drawing 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) in the top figures, we set ρ=0.3\rho=0.3. For this value of ρ\rho, s=0.6022s^{\star}=0.6022. Hence, we are in the feedback-limited regime for T=1.1,1.25T=1.1,1.25 and 1.51.5, and we observe exactly the same outage probabilities for both no-fading and fading cases.

On the other hand, we are in the rate-limited regime for T=2T=2, and the outage probabilities become the same for both selective feedback and all-feedback relay selections, i.e., see the red and black curves in the top figures. As a function of ρ\rho, the feedback-limited regime is manifested through the flat portion 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) in the bottom figures. In particular, when drawn as a function of ρ\rho, 𝖯out(𝒫FB,T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right) stays constant until a critical target rate, which is the feedback-limited regime. In this regime, the outage probability depends only on the average number of relays feeding back, i.e., 𝖯out(𝒫FB,T)=eμ(T){\sf P}_{\rm out}\left(\mathcal{P}_{\rm FB},T\right)={\rm e}^{-\mu\left(T\right)}. On the other hand, after a critical target rate value, outage probability curves coincide with each other and move together as a function of ρ\rho for both selective feedback and all-feedback scenarios. This is the rate-limited case, and the outage probability is independent of whether we employ a threshold-based selective feedback relay selection policy or not.

Based on our observations in Fig. 7 and earlier explanations after Theorem 5, as a practical network design rule of thumb, we can say that setting TT such that μ(T)5\mu\left(T\right)\approx 5 is enough to achieve the same average rate attained by the all-feedback policy with almost negligible performance loss. This point is specifically illustrated in Fig. 9 for both average rate and outage probability, where we observe almost no performance gap between the selective feedback and all-feedback approaches. However, it should be noted that an error floor phenomenon cannot be avoided with selective feedback for the outage probability performance as the network becomes denser due to the functional forms given in (51) and (55). For dense relay network deployments, the feedback load needs to be traded-off with outage probability to track the all-feedback outage performance.

Refer to caption
(a) Average rate vs λ\lambda for fixed feedback load.
Refer to caption
(b) Outage probability vs λ\lambda for fixed feedback load.
Figure 9: Average rate and outage probability achieved by the threshold-based selective feedback relay selection policy with a fixed feedback load at μ(T)=5\mu\left(T\right)=5.

IX Generalizations

IX-A Relays with FD Capability

In the FD case, the following modifications are required in (5) to obtain the corresponding data rates. We need to remove the scaling coefficient 12\frac{1}{2} in front of the minimum operator, scale the received signal power at the relay node with 1|ρ|21-\left|\rho\right|^{2}, add the term in (63)

FD Power Gain=2𝖲𝖭𝖱G(𝒙s𝒙d)G(𝒙𝒫𝒙d){ρHs,dHr,d}\displaystyle\mbox{FD Power Gain}=2{\sf SNR}\sqrt{G\left(\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}_{\rm d}\right\|\right)G\left(\left\|\boldsymbol{x}_{\mathcal{P}}-\boldsymbol{x}_{\rm d}\right\|\right)}\Re\left\{\rho H_{\rm s,d}H_{\rm r,d}^{*}\right\} (63)

to the received signal power at the destination node and optimize the correlation coefficient ρ\rho\in\mathbb{C} of the source and relay codebooks over the complex unit ball, where {x}\Re\left\{x\right\} and xx^{*} denote the real part and complex conjugate of xx\in\mathbb{C}, respectively, [9, 19, 45]. In the particular case of the severely shadowed source-destination links (i.e.,|Hs,d|0\left|H_{\rm s,d}\right|\approx 0) that we analyze in this paper, the FD power gain disappears and the FD rates are maximized when ρ=0\rho=0, which corresponds to independent codebook designs at the source and relay nodes.

Hence, for the severely shadowed source-destination links, the data rates achieved by the FD relay deployments are equivalent to those in the HD case given by (5), up to a scaling coefficient of 12\frac{1}{2}. This observation implies, for example, that the optimum relay location maximizing the HD rates continues to be the same for the FD operation. Therefore, our results obtained for the HD relays can be directly applied to this particular FD relay deployment scenario.

IX-B Non-homogeneous PPPs

Isotropy and complete randomness are two critical properties of HPPPs that enabled the derivation of key results, especially statistical characterization of Γopt\Gamma_{\rm opt} in Theorem 4, in the previous sections of the paper. In this part, we will derive the cdf and pdf of Γopt\Gamma_{\rm opt} for non-homogeneous but isotropic PPPs, which can be in turn used to obtain network performance measures such as 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) and 𝖯out(𝒫opt){\sf P}_{\rm out}\left(\mathcal{P}_{\rm opt}\right) for more general distributions of relays over 2\mathbb{R}^{2}. To this end, we consider a non-homogeneous PPP Φ\Phi having a mean measure Λ\Lambda, which is defined as

Λ(𝒮)𝖤[𝑿Φ𝟣{𝑿𝒮}]\displaystyle\Lambda\left(\mathcal{S}\right)\triangleq\mathsf{E}\left[\sum_{\boldsymbol{X}\in\Phi}\mathsf{1}_{\left\{\boldsymbol{X}\in\mathcal{S}\right\}}\right]

for any Borel subset 𝒮\mathcal{S} of 2\mathbb{R}^{2}. We say that Φ\Phi is an isotropic PPP if Λ\Lambda is invariant under rotations around the origin, i.e., Λ(𝒮)=Λ(Π(𝒮))\Lambda\left(\mathcal{S}\right)=\Lambda\left(\Pi\left(\mathcal{S}\right)\right) for all rotations Π\Pi around the origin and all Borel subsets 𝒮\mathcal{S} of 2\mathbb{R}^{2}. The next theorem provides the cdf and pdf of Γopt\Gamma_{\rm opt} for isotropic PPPs.

Theorem 8

For an isotropic PPP with mean measure Λ\Lambda, the cdf of Γopt\Gamma_{\rm opt} is given by

FΓopt(γ)={0 if γ<d1e2π0π2Λ((𝟎,γ2d2sin2θdcosθ))dθ if γd,\displaystyle F_{\Gamma_{\rm opt}}\left(\gamma\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ 1-{\rm e}^{-\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)\right){\rm d}\theta}&\mbox{ if }\gamma\geq d\end{array}\right., (66)

where (𝟎,r)\mathcal{B}\left(\boldsymbol{0},r\right) is the closed disc centered around the origin with radius rr. Further, if Λ\Lambda is absolutely continuous with respect to the Lebesgue measure with a continuous Radon-Nikodym derivative λ(𝐱)\lambda\left(\boldsymbol{x}\right), 𝐱2\boldsymbol{x}\in\mathbb{R}^{2}, then λ\lambda is a spherically symmetric function, i.e., λ(𝐱)=λ(𝐱)\lambda\left(\boldsymbol{x}\right)=\lambda\left(\left\|\boldsymbol{x}\right\|\right) for all 𝐱2\boldsymbol{x}\in\mathbb{R}^{2}, and the pdf of Γopt\Gamma_{\rm opt} is given by

fΓopt(γ)=40π2γ(γ2d2sin2θdcosθ)γ2d2sin2θλ(γ2d2sin2θdcosθ)dθ\displaystyle f_{\Gamma_{\rm opt}}\left(\gamma\right)=4\int_{0}^{\frac{\pi}{2}}\frac{\gamma\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)}{\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}}\lambda\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right){\rm d}\theta
e40π20γ2d2sin2θdcosθλ(ψ)ψdψdθ𝟣{γd}.\displaystyle\cdot{\rm e}^{-4\int_{0}^{\frac{\pi}{2}}\int_{0}^{\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta}\lambda\left(\psi\right)\psi{\rm d}\psi{\rm d}\theta}\cdot\mathsf{1}_{\left\{\gamma\geq d\right\}}. (67)
Proof:

See Appendix H. ∎

In the statement of Theorem 8, we represented the density λ\lambda, whenever it exists, as a function of a single variable due to its spherically symmetric nature with a slight abuse of notation. For an HPPP, Λ\Lambda is a scaled version of the Lebesgue measure, i.e., Λ(𝒮)=λ𝖺𝗋𝖾𝖺(𝒮)\Lambda\left(\mathcal{S}\right)=\lambda\cdot{\sf area}\left(\mathcal{S}\right) with constant λ\lambda, and it can be seen that Theorem 8 reduces to Theorem 4 after some manipulations. Below, we provide some examples, which are of either potential practical or theoretical interest, to illustrate the applications of Theorem 8. In all the examples below, we take 𝒙s=(d,0)\boldsymbol{x}_{\rm s}=\left(-d,0\right), 𝒙d=(d,0)\boldsymbol{x}_{\rm d}=\left(d,0\right).

Example 1: Φ\Phi is an HPPP with constant intensity λ>0\lambda>0 over 2(𝟎,r)\mathbb{R}^{2}\setminus\mathcal{B}\left(\boldsymbol{0},r\right) and Φ(𝟎,r)=\Phi\cap\mathcal{B}\left(\boldsymbol{0},r\right)=\emptyset. This is the situation in which there is a circular exclusion region with radius rr around the origin and none of the relays is allowed to lie in this region possibly due to an obstacle. The mean measure in this case is given by

Λ((𝟎,τ))={0 if τ<rλπ(τ2r2) if τr.\displaystyle\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\tau<r\\ \lambda\pi\left(\tau^{2}-r^{2}\right)&\mbox{ if }\tau\geq r\end{array}\right.. (70)

Using (66) and calculating the integral 0π2Λ((𝟎,γ2d2sin2θdcosθ))dθ\int_{0}^{\frac{\pi}{2}}\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)\right){\rm d}\theta for three different ranges of γ\gamma, we obtain

FΓopt(γ)={0 if γr2+d21e2λd2Id,r(γ)+2λr2arcsin(γ2d2r22dr) if r2+d2<γr+d1e2λd2((γd)2arcsec(γd)(γd)21)+λπr2 if γ>r+d,\displaystyle F_{\Gamma_{\rm opt}}\left(\gamma\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma\leq\sqrt{r^{2}+d^{2}}\\ 1-{\rm e}^{-2\lambda d^{2}I_{d,r}\left(\gamma\right)+2\lambda r^{2}\arcsin\left(\frac{\gamma^{2}-d^{2}-r^{2}}{2dr}\right)}&\mbox{ if }\sqrt{r^{2}+d^{2}}<\gamma\leq r+d\\ 1-{\rm e}^{-2\lambda d^{2}\left(\left(\frac{\gamma}{d}\right)^{2}\operatorname{arcsec}\left(\frac{\gamma}{d}\right)-\sqrt{\left(\frac{\gamma}{d}\right)^{2}-1}\right)+\lambda\pi r^{2}}&\mbox{ if }\gamma>r+d\end{array}\right., (74)

where Id,r(γ)I_{d,r}\left(\gamma\right) is given by (75) and (76).

Id,r(γ)=(γd)2(arcsec(γd)+arccsc(γd1bd,r(γ))arccos(γ2d2r22dr))\displaystyle I_{d,r}\left(\gamma\right)=\left(\frac{\gamma}{d}\right)^{2}\left(\operatorname{arcsec}\left(\frac{\gamma}{d}\right)+\operatorname{arccsc}\left(\frac{\gamma}{d}\frac{1}{b_{d,r}\left(\gamma\right)}\right)-\arccos\left(\frac{\gamma^{2}-d^{2}-r^{2}}{2dr}\right)\right)
+bd,r(γ)((γd)2bd,r2(γ)γ2d2r22dr)(γd)21\displaystyle+b_{d,r}\left(\gamma\right)\left(\sqrt{\left(\frac{\gamma}{d}\right)^{2}-b_{d,r}^{2}\left(\gamma\right)}-\frac{\gamma^{2}-d^{2}-r^{2}}{2dr}\right)-\sqrt{\left(\frac{\gamma}{d}\right)^{2}-1} (75)

 

bd,r(γ)=12dr(d+rγ)(d+r+γ)(γ+dr)(γ+rd)\displaystyle b_{d,r}\left(\gamma\right)=\frac{1}{2dr}\sqrt{\left(d+r-\gamma\right)\left(d+r+\gamma\right)\left(\gamma+d-r\right)\left(\gamma+r-d\right)} (76)

 

We note that Φ\Phi is an HPPP with intensity λ\lambda over 2\mathbb{R}^{2} when r=0r=0, and (74) reduces to (22) as expected. Since there is a circular exclusion region around the origin with radius rr, Γopt\Gamma_{\rm opt} always takes values larger than r2+d2\sqrt{r^{2}+d^{2}}, which results in the first case in (74). For γ>r+d\gamma>r+d, it can be seen that the function g(θ)=γ2d2sin2θdcosθg\left(\theta\right)=\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta is greater than rr for all θ[0,π2]\theta\in\left[0,\frac{\pi}{2}\right]. Hence, Λ((𝟎,g(θ)))=λπ(g2(θ)r2)\Lambda\left(\mathcal{B}\left(\boldsymbol{0},g\left(\theta\right)\right)\right)=\lambda\pi\left(g^{2}\left(\theta\right)-r^{2}\right) for all θ[0,π2]\theta\in\left[0,\frac{\pi}{2}\right], integration of which over [0,π2]\left[0,\frac{\pi}{2}\right] leads to the third case in (74). The functional form of the cdf of Γopt\Gamma_{\rm opt} in this case is identical to the one in (22), except an extra λπr2\lambda\pi r^{2} term due to the exclusion region around the origin. The most involved case is the one in which γ2+d2<γr+d\sqrt{\gamma^{2}+d^{2}}<\gamma\leq r+d. In this case, g(θ)rg\left(\theta\right)\leq r for θ[0,θ]\theta\in\left[0,\theta^{\star}\right] and g(θ)>rg\left(\theta\right)>r for θ(θ,π2]\theta\in\left.\left(\theta^{\star},\frac{\pi}{2}\right]\right., where θ=arccos(γ2d2r22dr)\theta^{\star}=\arccos\left(\frac{\gamma^{2}-d^{2}-r^{2}}{2dr}\right). Considering both integration intervals, we arrive at the second case in (74).

In the next two examples, we consider finite number of relays with a Poisson distribution, which will eventually lead to having defective Γopt\Gamma_{\rm opt} distributions.

Example 2: In this example, we consider the situation in which the relays form a HPPP Φ\Phi on a circle around the origin, which we represent as the boundary (𝟎,r)\partial\mathcal{B}\left(\boldsymbol{0},r\right) of (𝟎,r)\mathcal{B}\left(\boldsymbol{0},r\right). There is no other relay in the network. Then, the mean measure of Φ\Phi is given by

Λ((𝟎,τ))={0 if τ<r2λπr if τr.\displaystyle\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\tau<r\\ 2\lambda\pi r&\mbox{ if }\tau\geq r\end{array}\right.. (79)

When compared to the previously studied HPPP cases over 2\mathbb{R}^{2} with and without an exclusion region, one important feature of this case is that the total number of relays NN contained in Φ\Phi is a Poisson distributed random variable with mean 2λπr2\lambda\pi r, which takes finite values. On the other hand, Φ\Phi contained infinitely many relays in the previous cases. Using the master equation (66), we obtain the cdf of Γopt\Gamma_{\rm opt} for Poisson distributed relays on the circle as

FΓopt(γ)={0 if γr2+d21e4λrarcsin(γ2d2r22dr) if r2+d2<γr+d1e2λπr if γ>r+d.\displaystyle F_{\Gamma_{\rm opt}}\left(\gamma\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma\leq\sqrt{r^{2}+d^{2}}\\ 1-{\rm e}^{-4\lambda r\arcsin\left(\frac{\gamma^{2}-d^{2}-r^{2}}{2dr}\right)}&\mbox{ if }\sqrt{r^{2}+d^{2}}<\gamma\leq r+d\\ 1-{\rm e}^{-2\lambda\pi r}&\mbox{ if }\gamma>r+d\end{array}\right.. (83)

The first case in (83) reflects the fact that Γopt\Gamma_{\rm opt} is greater than r2+d2\sqrt{r^{2}+d^{2}} when relays are located on (𝟎,r)\partial\mathcal{B}\left(\boldsymbol{0},r\right). In the second case, g(θ)g(\theta), as defined in Example 1, is smaller than rr for θ[0,θ]\theta\in\left[0,\theta^{\star}\right] and larger than rr for θ(θ,π2]\theta\in\left.\left(\theta^{\star},\frac{\pi}{2}\right]\right., where θ=arccos(γ2d2r22dr)\theta^{\star}=\arccos\left(\frac{\gamma^{2}-d^{2}-r^{2}}{2dr}\right). For γ>r+d\gamma>r+d, we have g(θ)>rg\left(\theta\right)>r for all θ[0,π2]\theta\in\left[0,\frac{\pi}{2}\right] and all the relays available in the network are used to determine the value for FΓopt(γ)F_{\Gamma_{\rm opt}}\left(\gamma\right). However, since 𝖯𝗋{N=0}=e2λπr\mathsf{Pr}\left\{N=0\right\}={\rm e}^{-2\lambda\pi r}, some probability mass for Γopt\Gamma_{\rm opt} always escapes to infinity due to lack of a relay in the network that can provide connectivity between source and destination nodes.

Example 3: Φ\Phi is a Gaussian PPP with intensity function λ(𝒙)=n2πσ2e𝒙22σ2\lambda\left(\boldsymbol{x}\right)=\frac{n}{2\pi\sigma^{2}}{\rm e}^{-\frac{\left\|\boldsymbol{x}\right\|^{2}}{2\sigma^{2}}}, where nn is the total average number of relays contained in Φ\Phi [93]. This is the case in which relays are scattered around the origin according to a bell shaped density with exponentially decaying tails to provide connectivity between source and destination. For Gaussian PPP distribution of relays, the mean measure of Φ\Phi is equal to Λ((𝟎,τ))=n(1eτ22σ2)\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)=n\left(1-{\rm e}^{-\frac{\tau^{2}}{2\sigma^{2}}}\right) and the cdf of Γopt\Gamma_{\rm opt} can be obtained as

FΓopt(γ)={0 if γ<d1exp(n+2nπeγ22σ201ed2σ2(12u221u2(γd)2u2)1u2du) if γd\displaystyle F_{\Gamma_{\rm opt}}\left(\gamma\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ 1-\exp\left(-n+\frac{2n}{\pi}{\rm e}^{\frac{-\gamma^{2}}{2\sigma^{2}}\int_{0}^{1}\frac{{\rm e}^{\frac{-d}{2\sigma^{2}}\left(1-2u^{2}-2\sqrt{1-u^{2}}\sqrt{\left(\frac{\gamma}{d}\right)^{2}-u^{2}}\right)}}{\sqrt{1-u^{2}}}{\rm d}u}\right)&\mbox{ if }\gamma\geq d\end{array}\right. (86)

by using (66).

IX-C Different 𝖲𝖭𝖱{\sf SNR} Values at Relays and Destination

As another extension of our baseline model presented in the previous sections, we now consider a more heterogeneous communications scenario in which the 𝖲𝖭𝖱{\sf SNR} values at relays and the destination are different. For example, this is the operating situation of interest when transmission powers and/or RF circuitry at source and relay nodes are distinct. To model this situation, we will denote the 𝖲𝖭𝖱{\sf SNR} value at relays by 𝖲𝖭𝖱1{\sf SNR}_{1} and that at the destination node by 𝖲𝖭𝖱2{\sf SNR}_{2}. Conditioned on relay locations Φ\Phi, the rate achieved by a relay selection policy 𝒫\mathcal{P} is given according to

𝖱Φ(𝒫)=12min{𝖤[log2(1+𝖲𝖭𝖱1|Hs,r|2𝒙s𝑿𝒫α)|Φ],\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}\right)=\frac{1}{2}\min\bigl{\{}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}_{1}\left|H_{\rm s,r}\right|^{2}\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{X}_{\mathcal{P}}\right\|^{-\alpha}\right)\big{|}\Phi\right],
𝖤[log2(1+𝖲𝖭𝖱2|Hr,d|2𝑿𝒫𝒙dα)|Φ]},\displaystyle\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}_{2}\left|H_{\rm r,d}\right|^{2}\left\|\boldsymbol{X}_{\mathcal{P}}-\boldsymbol{x}_{\rm d}\right\|^{-\alpha}\right)\big{|}\Phi\right]\bigr{\}}, (87)

where we have taken G(x)=1xαG(x)=\frac{1}{x^{\alpha}}. Using (87) and following the arguments similar to those in Section IV, we can write the optimum relay selection problem as the minimization of the modified relay selection function s^diff(𝒙)\widehat{s}_{\rm diff}\left(\boldsymbol{x}\right), which is given by

s^diff(𝒙)=max{𝖲𝖭𝖱~1𝒙s𝒙,𝖲𝖭𝖱~2𝒙𝒙d},\displaystyle\widehat{s}_{\rm diff}\left(\boldsymbol{x}\right)=\max\left\{\widetilde{{\sf SNR}}_{1}\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}\right\|,\widetilde{{\sf SNR}}_{2}\left\|\boldsymbol{x}-\boldsymbol{x}_{\rm d}\right\|\right\}, (88)

where 𝖲𝖭𝖱~1=𝖲𝖭𝖱11/α\widetilde{{\sf SNR}}_{1}={\sf SNR}_{1}^{-1/\alpha} and 𝖲𝖭𝖱~2=𝖲𝖭𝖱21/α\widetilde{{\sf SNR}}_{2}={\sf SNR}_{2}^{-1/\alpha}. Here, we can interpret 𝖲𝖭𝖱~1\widetilde{{\sf SNR}}_{1} and 𝖲𝖭𝖱~2\widetilde{{\sf SNR}}_{2} as effective 𝖲𝖭𝖱{\sf SNR}s at relays and destination, respectively, for optimum relay selection. The CQI at the optimum relay node in Φ\Phi is now given according to

Γopt,diff=min𝑿Φs^diff(𝑿).\displaystyle\Gamma_{\rm opt,diff}=\min_{\boldsymbol{X}\in\Phi}\widehat{s}_{\rm diff}\left(\boldsymbol{X}\right). (89)

Accordingly, the conditional data rate achieved by the optimum relay selection policy 𝒫opt\mathcal{P}_{\rm opt}, given the relay locations, is equal to

𝖱Φ(𝒫opt)=12𝖤[log2(1+|H|2Γopt,diffα)|Φ].\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right)=\frac{1}{2}\mathsf{E}\left[\log_{2}\left(1+\left|H\right|^{2}\Gamma_{\rm opt,diff}^{-\alpha}\right)\big{|}\Phi\right]. (90)

As in the homogeneous case, the expression in (90) shows that it is enough to derive the distribution of Γopt,diff\Gamma_{\rm opt,diff} in order to carry out the same analysis for the average rate and outage probability as is done in Section VI. The cdf of Γopt,diff\Gamma_{\rm opt,diff} is obtained in the following theorem.

Theorem 9

The cdf of Γopt,diff\Gamma_{\rm opt,diff} is given by

FΓopt,diff(γ)={0 if γ<2d(𝖲𝖭𝖱~1𝖲𝖭𝖱~2𝖲𝖭𝖱~1+𝖲𝖭𝖱~2)1eλ(γ𝖲𝖭𝖱~1)2sec1(4d(γ𝖲𝖭𝖱~1)(γ𝖲𝖭𝖱~1)2(γ𝖲𝖭𝖱~2)2+4d2)eλ(γ𝖲𝖭𝖱~2)2sec1(4d(γ𝖲𝖭𝖱~2)(γ𝖲𝖭𝖱~2)2(γ𝖲𝖭𝖱~1)2+4d2)eλ(γ𝖲𝖭𝖱~1+γ𝖲𝖭𝖱~22d)(γ𝖲𝖭𝖱~1γ𝖲𝖭𝖱~2+2d)(γ𝖲𝖭𝖱~1+γ𝖲𝖭𝖱~2+2d)(γ𝖲𝖭𝖱~1+γ𝖲𝖭𝖱~2+2d) if γ2d(𝖲𝖭𝖱~1𝖲𝖭𝖱~2𝖲𝖭𝖱~1+𝖲𝖭𝖱~2).\begin{split}F_{\Gamma_{\rm opt,diff}}\left(\gamma\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<2d\left(\frac{\widetilde{\mathsf{SNR}}_{1}\widetilde{\mathsf{SNR}}_{2}}{\widetilde{\mathsf{SNR}}_{1}+\widetilde{\mathsf{SNR}}_{2}}\right)\\ 1-{\rm e}^{-\lambda\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}\right)^{2}\sec^{-1}\left(\frac{4d\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}\right)}{\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}\right)^{2}-\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}\right)^{2}+4d^{2}}\right)}\\ \qquad{\rm e}^{-\lambda\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}\right)^{2}\sec^{-1}\left(\frac{4d\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}\right)}{\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}\right)^{2}-\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}\right)^{2}+4d^{2}}\right)}\\ \qquad{\rm e}^{\lambda\sqrt{\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}+\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}-2d\right)\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}-\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}+2d\right)\left(-\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}+\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}+2d\right)\left(\frac{\gamma}{\widetilde{\mathsf{SNR}}_{1}}+\frac{\gamma}{\widetilde{\mathsf{SNR}}_{2}}+2d\right)}}\\ &\mbox{ if }\gamma\geq 2d\left(\frac{\widetilde{\mathsf{SNR}}_{1}\widetilde{\mathsf{SNR}}_{2}}{\widetilde{\mathsf{SNR}}_{1}+\widetilde{\mathsf{SNR}}_{2}}\right)\end{array}\right..\end{split} (91)
Proof:

See Appendix I. ∎

An important remark about the result in Theorem 9 is that the optimum relay location does not satisfy the distance balancing property anymore due to different 𝖲𝖭𝖱{\sf SNR} scalings at the source-relay and relay-destination links. In particular, the minimum of s^diff(𝒙)\widehat{s}_{\rm diff}\left(\boldsymbol{x}\right) is achieved at a point 𝒙\boldsymbol{x}^{\star} located on the line segment connecting 𝒙s\boldsymbol{x}_{\rm s} and 𝒙d\boldsymbol{x}_{\rm d} that satisfies 𝒙𝒙d=2d𝖲𝖭𝖱~1𝖲𝖭𝖱~1+𝖲𝖭𝖱~2\left\|\boldsymbol{x}^{\star}-\boldsymbol{x}_{\rm d}\right\|=2d\frac{\widetilde{\mathsf{SNR}}_{1}}{\widetilde{\mathsf{SNR}}_{1}+\widetilde{\mathsf{SNR}}_{2}}. Hence, the relays in Φ\Phi closer to 𝒙\boldsymbol{x}^{\star} are likely to be the better candidates to pick to maximize the system performance when the relays and destination are subject to different noise conditions.

Refer to caption
Figure 10: The cdf of Γopt,diff\Gamma_{\rm opt,diff} for different 𝖲𝖭𝖱{\sf SNR}s at relay and destination nodes when d=0.5d=0.5 and λ=1\lambda=1.

X Conclusions

In this paper, we have studied a relay-aided wireless network with a single source-destination pair and spatially deployed decode-and-forward relays. We have obtained structural properties for the optimum relay selection policy and the distribution of the channel quality indicator of the relay node selected by the optimum policy. To benchmark the optimum relay selection scheme, we have analyzed the mid-point relay selection policy and obtained a sufficient condition for its optimality. These results hold for general fading distributions and non-increasing path-loss models decaying to zero. Using the derived distribution of the optimum channel quality indicator, we have also characterized best achievable average rates and minimum outage probability for Rayleigh fading channels and for when there is no fading.

An important practical limitation to implement optimum relay selection policy in physical wireless systems is the feedback load required during the relay selection process. To alleviate this limitation, we have proposed a threshold-based distributed relay selection strategy in which each relay node gives an autonomous feedback decision based on its local channel quality indicator. For this class of relay selection policies, we have shown that the total number of relays feeding back is a Poisson distributed random variable. We have characterized the average value for this Poisson distribution analytically, and obtained the analytical expressions for the average rate and outage probability achieved with reduced feedback load for both the no-fading and Rayleigh fading communications scenarios. We have derived useful and practical design rules for selecting relay nodes in a distributed way, which indicates that setting the threshold value to have five relay nodes feeding back on average is enough to achieve almost the same communications performance attained by the all-feedback policy with negligible performance loss. The performance loss becomes insignificant especially when the relay intensity increases.

Utilizing the developed analytical framework, an important future plan of the authors is the generalization of the derived results to three-dimensional network topologies. A significant application scenario of this generalization is unmanned-aerial-vehicle (UAV) aided wireless communications. The deployment and trajectory optimization of UAVs for coverage and capacity augmentation of terrestrial networks have received increasing interest in recent years [94, 95]. This extension will involve a study of the probabilistic path-loss functions for ground-to-air and air-to-ground channels as well as a study of stochastic geometry models on three dimensional manifolds to integrate the UAV specific communications aspects with the developed framework.

Appendix A Proof of Theorem 2

Without loss of generality, we take 𝒘=𝟎\boldsymbol{w}=\boldsymbol{0} and let suff={s^(𝑿mid)d2+𝑿(2)2}\mathcal{E}_{\rm suff}=\left\{\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\leq\sqrt{d^{2}+\left\|\boldsymbol{X}_{(2)}\right\|^{2}}\right\}, where 𝑿midΦ\boldsymbol{X}_{\rm mid}\in\Phi and 𝑿(2)Φ\boldsymbol{X}_{(2)}\in\Phi are the closest and second closest points of Φ\Phi to 𝟎\boldsymbol{0}. Let Ψmid=𝑿mid\Psi_{\rm mid}=\left\|\boldsymbol{X}_{\rm mid}\right\|, Θmid\Theta_{\rm mid} be the angle of 𝑿mid\boldsymbol{X}_{\rm mid} and Ψ(2)=𝑿(2)\Psi_{(2)}=\left\|\boldsymbol{X}_{(2)}\right\|. Let also fNN(ψ)f_{\rm NN}\left(\psi\right) be the pdf of Ψmid\Psi_{\rm mid}, which is the nearest-neighbor distribution for HPPPs. Due to isotropy of HPPPs, Θmid\Theta_{\rm mid} and Ψ(2)\Psi_{(2)} are independent random variables and Θmid\Theta_{\rm mid} is uniformly distributed over [0,2π)\left.\left[0,2\pi\right)\right.. Using these observations, 𝖯𝗋(suff)\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) can be expressed as

𝖯𝗋(suff)\displaystyle\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) =\displaystyle= 2π0π20𝖯𝗋(suff|Ψmid=ψ,Θmid=θ)fNN(ψ)dψdθ\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\int_{0}^{\infty}\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\ \big{|}\ \Psi_{\rm mid}=\psi,\Theta_{\rm mid}=\theta\right)f_{\rm NN}\left(\psi\right){\rm d}\psi{\rm d}\theta (92)
=\displaystyle= 2π0π20𝖯𝗋{Ψ(2)ψ2+2dψcosθ|Ψmid=ψ}fNN(ψ)dψdθ.\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\int_{0}^{\infty}\mathsf{Pr}\left\{\Psi_{(2)}\geq\sqrt{\psi^{2}+2d\psi\cos\theta}\ \Big{|}\ \Psi_{\rm mid}=\psi\right\}f_{\rm NN}\left(\psi\right){\rm d}\psi{\rm d}\theta.

Using spherical contact distributions for HPPPs [93], it can be shown that

𝖯𝗋{Ψ(2)<ψ|Ψmid=ψ}={0 if ψ<ψ1eλπ((ψ)2ψ2) if ψψ.\mathsf{Pr}\left\{\Psi_{(2)}<\psi^{\prime}\big{|}\Psi_{\rm mid}=\psi\right\}=\left\{\begin{array}[]{ll}0&\mbox{ if }\psi^{\prime}<\psi\\ 1-{\rm e}^{-\lambda\pi\left(\left(\psi^{\prime}\right)^{2}-\psi^{2}\right)}&\mbox{ if }\psi^{\prime}\geq\psi\end{array}\right.. (93)

Using (92) and (93), we have

𝖯𝗋(suff)\displaystyle\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) =\displaystyle= 2π0π20e2λπdcos(θ)ψfNN(ψ)dψdθ\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\int_{0}^{\infty}{\rm e}^{-2\lambda\pi d\cos(\theta)\psi}f_{\rm NN}(\psi){\rm d}\psi{\rm d}\theta (94)
=\displaystyle= 2π0π2Ψmid(2λπdcosθ)dθ,\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\mathcal{M}_{\Psi_{\rm mid}}(-2\lambda\pi d\cos\theta){\rm d}\theta,

where Ψmid(t)=𝖤[etΨmid]\mathcal{M}_{\Psi_{\rm mid}}(t)=\mathsf{E}\left[e^{t\cdot\Psi_{\rm mid}}\right] is the moment generating function of Ψmid\Psi_{\rm mid}. Since Ψmid\Psi_{\rm mid} is Rayleigh distributed with pdf fNN(ψ)=2λπψeλπψ2f_{\rm NN}\left(\psi\right)=2\lambda\pi\psi{\rm e}^{-\lambda\pi\psi^{2}} [76], Ψmid(t)\mathcal{M}_{\Psi_{\rm mid}}(t) is given by

Ψmid(t)=1+t2λ(erf(t2λπ)+1)et24λπ,\displaystyle\mathcal{M}_{\Psi_{\rm mid}}(t)=1+\frac{t}{2\sqrt{\lambda}}\left({\rm erf}\left(\frac{t}{2\sqrt{\lambda\pi}}\right)+1\right){\rm e}^{\frac{t^{2}}{4\lambda\pi}},

where erf(){\rm erf}\left(\cdot\right) is the Gauss error function. As a result, we can write 𝖯𝗋(suff)\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) in (94) as

𝖯𝗋(suff)\displaystyle\mathsf{Pr}\left(\mathcal{E}_{\rm suff}\right) =\displaystyle= 2π0π2(1λπdcos(θ)eλπd2cos2(θ)erfc(λπdcos(θ)))dθ\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\left(1-\sqrt{\lambda}\pi d\cos(\theta){\rm e}^{\lambda\pi d^{2}\cos^{2}(\theta)}{\rm erfc}\left(\sqrt{\lambda\pi}d\cos(\theta)\right)\right){\rm d}\theta (95)
=(a)\displaystyle\stackrel{{\scriptstyle\rm(a)}}{{=}} eλπd2erfc(λπd),\displaystyle{\rm e}^{\lambda\pi d^{2}}{\rm erfc}\left(\sqrt{\lambda\pi}d\right),

where erfc(x)=1erf(x){\rm erfc}\left(x\right)=1-{\rm erf}\left(x\right) is the complementary error function, and (a) follows as we first use the integral identity [96, eq. 3.2.6]

0ea2tt+cos2θdt=πaea2cos2θerfc(acosθ),\int_{0}^{\infty}\frac{{\rm e}^{-a^{2}t}}{\sqrt{t+\cos^{2}\theta}}{\rm d}t=\frac{\sqrt{\pi}}{a}{\rm e}^{a^{2}\cos^{2}\theta}{\rm erfc}\left(a\cos\theta\right),

then apply the integral representation [96, eq. 3.2.7]

erfc(x)=2πex20ex2t2t2+1dt,{\rm erfc}\left(x\right)=\frac{2}{\pi}{\rm e}^{-x^{2}}\int_{0}^{\infty}\frac{{\rm e}^{-x^{2}t^{2}}}{t^{2}+1}{\rm d}t,

and with some further straight forward mathematical manipulations.

Appendix B Proof of Lemma 4

In this appendix, we provide a key lemma that will be used to prove Theorem 3 and Theorem 5. For τψ0\tau\geq\psi\geq 0, let 𝒟(ψ,τ)=(𝟎,τ)(𝟎,ψ)\mathcal{D}(\psi,\tau)=\mathcal{B}\left(\boldsymbol{0},\tau\right)\setminus\mathcal{B}\left(\boldsymbol{0},\psi\right) with (𝟎,r)\mathcal{B}\left(\boldsymbol{0},r\right) being the closed disc centered around the origin 𝟎\boldsymbol{0} with radius rr, and 𝑼\boldsymbol{U} be uniformly distributed over 𝒟(ψ,τ)\mathcal{D}(\psi,\tau). The following lemma provides an expression for 𝖯𝗋{s^(𝑼)>t}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\}.

Lemma 4

Let 𝐱s=(d,0)\boldsymbol{x}_{\rm s}=\left(-d,0\right)^{\top} and 𝐱d=(d,0)\boldsymbol{x}_{\rm d}=\left(d,0\right)^{\top}. For τψ2+2dψ\tau\geq\sqrt{\psi^{2}+2d\psi}, 𝖯𝗋{s^(𝐔)>t}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} is given by

𝖯𝗋{s^(𝑼)>t}={1 if t<ψ2+d2τ2t2τ2ψ2+2a(t2ψ2)+d2sin(2a)π(τ2ψ2)+pτ,ψ(t,d)pτ,ψ(t,dsin(a)) if ψ2+d2t<ψ+dτ2t2τ2ψ2+pτ,ψ(t,d) if ψ+dt<τ2+d22b(τ2t2)d2sin(2b)π(τ2ψ2)+pτ,ψ(t,dsin(b)) if τ2+d2t<τ+d0 if tτ+d,\begin{split}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\}&=\left\{\begin{array}[]{ll}1&\text{ if }t<\sqrt{\psi^{2}+d^{2}}\\ \frac{\tau^{2}-t^{2}}{\tau^{2}-\psi^{2}}+\frac{2a^{\star}(t^{2}-\psi^{2})+d^{2}\sin(2a^{\star})}{\pi\left(\tau^{2}-\psi^{2}\right)}\\ \hskip 29.87547pt+p_{\tau,\psi}(t,d)-p_{\tau,\psi}(t,d\sin(a^{\star}))&\text{ if }\sqrt{\psi^{2}+d^{2}}\leq t<\psi+d\\ \frac{\tau^{2}-t^{2}}{\tau^{2}-\psi^{2}}+p_{\tau,\psi}(t,d)&\text{ if }\psi+d\leq t<\sqrt{\tau^{2}+d^{2}}\\ \frac{2b^{\star}(\tau^{2}-t^{2})-d^{2}\sin(2b^{\star})}{\pi(\tau^{2}-\psi^{2})}+p_{\tau,\psi}(t,d\sin(b^{\star}))&\text{ if }\sqrt{\tau^{2}+d^{2}}\leq t<\tau+d\\ 0&\text{ if }t\geq\tau+d\end{array}\right.,\end{split}

where pτ,ψ(t,d)2dt2d2π(τ2ψ2)+2t2π(τ2ψ2)arctan(dt2d2)p_{\tau,\psi}(t,d)\triangleq\frac{2d\sqrt{t^{2}-d^{2}}}{\pi(\tau^{2}-\psi^{2})}+\frac{2t^{2}}{\pi(\tau^{2}-\psi^{2})}\arctan\left(\frac{d}{\sqrt{t^{2}-d^{2}}}\right), a=arccos(t2ψ2d22dψ)a^{\star}=\arccos\left(\frac{t^{2}-\psi^{2}-d^{2}}{2d\psi}\right) and b=arccos(t2τ2d22dτ)b^{\star}=\arccos\left(\frac{t^{2}-\tau^{2}-d^{2}}{2d\tau}\right).

Proof:

The lemma directly follows for t<ψ2+d2t<\sqrt{\psi^{2}+d^{2}} and tτ+dt\geq\tau+d since ψ2+d2s^(𝒙)τ+d\sqrt{\psi^{2}+d^{2}}\leq\widehat{s}\left(\boldsymbol{x}\right)\leq\tau+d for all 𝒙𝒟(ψ,τ)\boldsymbol{x}\in\mathcal{D}(\psi,\tau). Hence, we will only focus on the case ψ2+d2t<τ+d\sqrt{\psi^{2}+d^{2}}\leq t<\tau+d. Let Ψ=𝑼\Psi=\left\|\boldsymbol{U}\right\| and Θ\Theta be the angle of 𝑼\boldsymbol{U}. Ψ\Psi and Θ\Theta are independent random variables due to isotropy, having pdfs fΨ(ψ)=2ψτ2ψ2f_{\Psi}(\psi^{\prime})=\frac{2\psi^{\prime}}{\tau^{2}-\psi^{2}} for ψψτ\psi\leq\psi^{\prime}\leq\tau and fΘ(θ)=12πf_{\Theta}(\theta)=\frac{1}{2\pi} for θ[0,2π)\theta\in[0,2\pi), respectively. By conditioning on Θ\Theta, we can express 𝖯𝗋{s^(𝑼)>t}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} as

𝖯𝗋{s^(𝑼)>t}\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} =\displaystyle= 12π02π𝖯𝗋{Ψ2+2d|cos(θ)|Ψ+d2>t2}dθ\displaystyle\frac{1}{2\pi}\int_{0}^{2\pi}\mathsf{Pr}\left\{\Psi^{2}+2d\left|\cos\left(\theta\right)\right|\Psi+d^{2}>t^{2}\right\}{\rm d}\theta (96)
=\displaystyle= 12π02π𝖯𝗋{Ψ>d|cos(θ)|+t2d2sin2(θ)}dθ\displaystyle\frac{1}{2\pi}\int_{0}^{2\pi}\mathsf{Pr}\left\{\Psi>-d\left|\cos\left(\theta\right)\right|+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right\}{\rm d}\theta
=\displaystyle= 2π0π2𝖯𝗋{Ψ>dcos(θ)+t2d2sin2(θ)}dθ,\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\mathsf{Pr}\left\{\Psi>-d\cos\left(\theta\right)+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right\}{\rm d}\theta,

where the last equality follows from the symmetry in the problem.

Let g(θ,t)𝖯𝗋{Ψ>d|cos(θ)|+t2d2sin2(θ)}g\left(\theta,t\right)\triangleq\mathsf{Pr}\left\{\Psi>-d\left|\cos\left(\theta\right)\right|+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right\}, tmin(θ)ψ2+2d|cos(θ)|ψ+d2t_{\min}\left(\theta\right)\triangleq\sqrt{\psi^{2}+2d\left|\cos\left(\theta\right)\right|\psi+d^{2}} and tmax(θ)τ2+2d|cos(θ)|τ+d2t_{\max}\left(\theta\right)\triangleq\sqrt{\tau^{2}+2d\left|\cos\left(\theta\right)\right|\tau+d^{2}}. It can be seen that g(θ,t)=1g\left(\theta,t\right)=1 if t<tmin(θ)t<t_{\min}\left(\theta\right) and g(θ,t)=0g\left(\theta,t\right)=0 if t>tmax(θ)t>t_{\max}\left(\theta\right). Hence, g(θ,t)g\left(\theta,t\right) can be written as

g(θ,t)=𝟣{t<tmin(θ)}+τ2(d|cos(θ)|+t2d2sin2(θ))2τ2ψ2𝟣{tmin(θ)ttmax(θ)}.\displaystyle g\left(\theta,t\right)=\mathsf{1}_{\left\{t<t_{\min}\left(\theta\right)\right\}}+\frac{\tau^{2}-\left(-d\left|\cos\left(\theta\right)\right|+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right)^{2}}{\tau^{2}-\psi^{2}}\mathsf{1}_{\left\{t_{\min}\left(\theta\right)\leq t\leq t_{\max}\left(\theta\right)\right\}}. (97)

We will consider three disjoint intervals for tt to calculate 𝖯𝗋{s^(𝑼)>t}=2π0π2g(θ,t)𝑑θ\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\}=\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}g\left(\theta,t\right)d\theta by using (97). We first consider ψ2+d2t<ψ+d\sqrt{\psi^{2}+d^{2}}\leq t<\psi+d. We note that tmin(θ)t_{\min}\left(\theta\right) is a continuous and strictly decreasing function of θ\theta for θ[0,π2]\theta\in\left[0,\frac{\pi}{2}\right]. Further, tmin(0)=ψ+dt_{\min}\left(0\right)=\psi+d and tmin(π2)=ψ2+d2t_{\min}\left(\frac{\pi}{2}\right)=\sqrt{\psi^{2}+d^{2}}. Hence, t<tmin(θ)t<t_{\min}\left(\theta\right) for θ[0,a)\theta\in\left.\left[0,a^{\star}\right)\right., where a=arccos(t2ψ2d22dψ)a^{\star}=\arccos\left(\frac{t^{2}-\psi^{2}-d^{2}}{2d\psi}\right). For θ[a,π2]\theta\in\left[a^{\star},\frac{\pi}{2}\right], we have tmin(θ)ttmax(θ)t_{\min}\left(\theta\right)\leq t\leq t_{\max}\left(\theta\right) since tmax(θ)t_{\max}\left(\theta\right) is a continuous and strictly decreasing function of θ\theta for θ[0,π2]\theta\in\left[0,\frac{\pi}{2}\right], having minimum value tmax(π2)=τ2+d2t_{\max}\left(\frac{\pi}{2}\right)=\sqrt{\tau^{2}+d^{2}} and τψ2+2dψ\tau\geq\sqrt{\psi^{2}+2d\psi}. Using these observations, we can write 𝖯𝗋{s^(𝑼)>t}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} as

𝖯𝗋{s^(𝑼)>t}\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} =\displaystyle= 2π0a𝟣{t<tmin(θ)}dθ+2πaπ2τ2(dcos(θ)+t2d2sin2(θ))2τ2ψ2dθ\displaystyle\frac{2}{\pi}\int_{0}^{a^{\star}}\mathsf{1}_{\left\{t<t_{\min}\left(\theta\right)\right\}}{\rm d}\theta+\frac{2}{\pi}\int_{a^{\star}}^{\frac{\pi}{2}}\frac{\tau^{2}-\left(-d\cos\left(\theta\right)+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right)^{2}}{\tau^{2}-\psi^{2}}{\rm d}\theta
=\displaystyle= 2πa+2πaπ2τ2(dcos(θ)+t2d2sin2(θ))2τ2ψ2dθ\displaystyle\frac{2}{\pi}a^{\star}+\frac{2}{\pi}\int_{a^{\star}}^{\frac{\pi}{2}}\frac{\tau^{2}-\left(-d\cos\left(\theta\right)+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right)^{2}}{\tau^{2}-\psi^{2}}{\rm d}\theta
=\displaystyle= τ2t2τ2ψ2+2a(t2ψ2)+d2sin(2a)π(τ2ψ2)+pτ,ψ(t,d)pτ,ψ(t,dsin(a))\displaystyle\frac{\tau^{2}-t^{2}}{\tau^{2}-\psi^{2}}+\frac{2a^{\star}\left(t^{2}-\psi^{2}\right)+d^{2}\sin\left(2a^{\star}\right)}{\pi(\tau^{2}-\psi^{2})}+p_{\tau,\psi}(t,d)-p_{\tau,\psi}\left(t,d\sin\left(a^{\star}\right)\right)

for ψ2+d2t<ψ+d\sqrt{\psi^{2}+d^{2}}\leq t<\psi+d.

Next, we consider ψ+dt<τ2+d2\psi+d\leq t<\sqrt{\tau^{2}+d^{2}}. In this case, we have tmin(θ)ttmax(θ)t_{\min}\left(\theta\right)\leq t\leq t_{\max}\left(\theta\right) for all θ[0,π2]\theta\in\left[0,\frac{\pi}{2}\right]. Hence,

𝖯𝗋{s^(𝑼)>t}\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} =\displaystyle= 2π0π2τ2(dcos(θ)+t2d2sin2(θ))2τ2ψ2dθ\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\frac{\tau^{2}-\left(-d\cos\left(\theta\right)+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right)^{2}}{\tau^{2}-\psi^{2}}{\rm d}\theta
=\displaystyle= τ2t2τ2ψ2+pτ,ψ(t,d)\displaystyle\frac{\tau^{2}-t^{2}}{\tau^{2}-\psi^{2}}+p_{\tau,\psi}\left(t,d\right)

for ψ+dt<τ2+d2\psi+d\leq t<\sqrt{\tau^{2}+d^{2}}. Finally, for τ2+d2t<τ+d\sqrt{\tau^{2}+d^{2}}\leq t<\tau+d, we have ttmax(θ)t\leq t_{\max}\left(\theta\right) for θ[0,b]\theta\in\left[0,b^{\star}\right] and t>tmax(θ)t>t_{\max}\left(\theta\right) for θ(b,π2]\theta\in\left.\left(b^{\star},\frac{\pi}{2}\right]\right., where b=arccos(t2τ2d22dτ)b^{\star}=\arccos\left(\frac{t^{2}-\tau^{2}-d^{2}}{2d\tau}\right). Thus,

𝖯𝗋{s^(𝑼)>t}\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} =\displaystyle= 2π0bτ2(dcos(θ)+t2d2sin2(θ))2τ2ψ2dθ\displaystyle\frac{2}{\pi}\int_{0}^{b^{\star}}\frac{\tau^{2}-\left(-d\cos\left(\theta\right)+\sqrt{t^{2}-d^{2}\sin^{2}\left(\theta\right)}\right)^{2}}{\tau^{2}-\psi^{2}}{\rm d}\theta
=\displaystyle= 2b(τ2t2)d2sin(2b)π(τ2ψ2)+pτ,ψ(t,dsin(b)),\displaystyle\frac{2b^{\star}\left(\tau^{2}-t^{2}\right)-d^{2}\sin\left(2b^{\star}\right)}{\pi\left(\tau^{2}-\psi^{2}\right)}+p_{\tau,\psi}\left(t,d\sin\left(b^{\star}\right)\right),

which concludes the proof. ∎

For the sake of completeness, we illustrate the analytical expression derived for 𝖯𝗋{s^(𝑼)>t}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>t\right\} in Lemma 4 and our simulation results for d=1d=1 and different values of τ\tau and ψ\psi parameters in Fig. 11. As can be seen in this figure, the analytical and simulation results perfectly match each other.

Refer to captionRefer to captionRefer to captionRefer to caption
Figure 11: Complementary cdf of s^(𝑼)\widehat{s}\left(\boldsymbol{U}\right) for d=1d=1 and different values of τ\tau and ψ\psi parameters. τ=10\tau=10 and ψ=2\psi=2 for upper left hand-side figure; τ=20\tau=20 and ψ=2\psi=2 for upper right hand-side figure; τ=10\tau=10 and ψ=5\psi=5 for bottom left hand-side figure; τ=20\tau=20 and ψ=5\psi=5 for bottom right hand-side figure.

Appendix C Proof of Theorem 3

Without loss of generality, we assume that 𝒙s=(d,0)\boldsymbol{x}_{\rm s}=\left(-d,0\right)^{\top} and 𝒙d=(d,0)\boldsymbol{x}_{\rm d}=\left(d,0\right)^{\top} due to stationarity and isotropy of HPPPs. Observing that s^(𝑿)s^(𝒀)\widehat{s}\left(\boldsymbol{X}\right)\neq\widehat{s}\left(\boldsymbol{Y}\right) for any two different points 𝑿\boldsymbol{X} and 𝒀\boldsymbol{Y} in Φ\Phi with probability one, we can write 𝖯𝗋{𝑿mid=𝑿opt}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} as

𝖯𝗋{𝑿mid=𝑿opt}\displaystyle\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} =\displaystyle= 𝖯𝗋{s^(𝑿mid)=s^(𝑿opt)}\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)=\widehat{s}\left(\boldsymbol{X}_{\rm opt}\right)\right\}
=\displaystyle= 𝖤𝑿mid[𝖯𝗋{s^(𝑿mid)<min𝑿Φ{𝑿mid}s^(𝑿)|𝑿mid}].\displaystyle\mathsf{E}_{\boldsymbol{X}_{\rm mid}}\left[\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)<\min_{\boldsymbol{X}\in\Phi\setminus\left\{\boldsymbol{X}_{\rm mid}\right\}}\widehat{s}\left(\boldsymbol{X}\right)\Big{|}\boldsymbol{X}_{\rm mid}\right\}\right].

Consider the function g(𝒙)=𝖯𝗋{s^(𝒙)<min𝑿Φ{𝒙}s^(𝑿)|𝑿mid=𝒙}g\left(\boldsymbol{x}\right)=\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{x}\right)<\min_{\boldsymbol{X}\in\Phi\setminus\left\{\boldsymbol{x}\right\}}\widehat{s}\left(\boldsymbol{X}\right)\Big{|}\boldsymbol{X}_{\rm mid}=\boldsymbol{x}\right\}. Let ψ=𝒙\psi=\left\|\boldsymbol{x}\right\| and (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right) be the closed disc centered around the origin and having radius τ=ψ2+2dψ\tau=\sqrt{\psi^{2}+2d\psi}. For any 𝒚2(𝟎,τ)\boldsymbol{y}\in\mathbb{R}^{2}\setminus\mathcal{B}\left(\boldsymbol{0},\tau\right), we have s^(𝒚)>ψ+d\widehat{s}\left(\boldsymbol{y}\right)>\psi+d and s^(𝒙)ψ+d\widehat{s}\left(\boldsymbol{x}\right)\leq\psi+d. Hence, we can express g(𝒙)g\left(\boldsymbol{x}\right) as

g(𝒙)\displaystyle g\left(\boldsymbol{x}\right) =\displaystyle= 𝖯𝗋{s^(𝒙)<min𝑿Φ(𝟎,τ){𝒙}s^(𝑿)|𝑿mid=𝒙}\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{x}\right)<\min_{\boldsymbol{X}\in\Phi\cap\mathcal{B}\left(\boldsymbol{0},\tau\right)\setminus\left\{\boldsymbol{x}\right\}}\widehat{s}\left(\boldsymbol{X}\right)\Big{|}\boldsymbol{X}_{\rm mid}=\boldsymbol{x}\right\} (98)
=\displaystyle= 𝖯𝗋{s^(𝒙)<min𝑿Φ𝒟(ψ,τ)s^(𝑿)|𝑿mid=𝒙},\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{x}\right)<\min_{\boldsymbol{X}\in\Phi\cap\mathcal{D}\left(\psi,\tau\right)}\widehat{s}\left(\boldsymbol{X}\right)\Big{|}\boldsymbol{X}_{\rm mid}=\boldsymbol{x}\right\},
=\displaystyle= 𝖯𝗋{s^(𝒙)<min𝑿Φ𝒟(ψ,τ)s^(𝑿)}\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{x}\right)<\min_{\boldsymbol{X}\in\Phi\cap\mathcal{D}\left(\psi,\tau\right)}\widehat{s}\left(\boldsymbol{X}\right)\right\}

where 𝒟(ψ,τ)=(𝟎,τ)(𝟎,ψ)\mathcal{D}\left(\psi,\tau\right)=\mathcal{B}\left(\boldsymbol{0},\tau\right)\setminus\mathcal{B}\left(\boldsymbol{0},\psi\right), the second equality follows from the fact 𝑿mid\boldsymbol{X}_{\rm mid} is at the boundary of (𝟎,ψ)\mathcal{B}\left(\boldsymbol{0},\psi\right) given 𝑿mid=𝒙\boldsymbol{X}_{\rm mid}=\boldsymbol{x} and there is no relay in the interior of (𝟎,ψ)\mathcal{B}\left(\boldsymbol{0},\psi\right), and the last equality follows from the total independence property of HPPPs. Below, we will obtain the functional form of g(𝒙)g\left(\boldsymbol{x}\right) to complete the proof.

To this end, let NN be the number of relays in Φ𝒟(ψ,τ)\Phi\cap\mathcal{D}\left(\psi,\tau\right). Given the event {N=n}\left\{N=n\right\} for n1n\geq 1, all the relays in Φ𝒟(ψ,τ)\Phi\cap\mathcal{D}\left(\psi,\tau\right) are uniformly distributed over 𝒟(ψ,τ)\mathcal{D}\left(\psi,\tau\right). Therefore, Lemma 4 can be used directly to obtain the functional form for g(𝒙)g\left(\boldsymbol{x}\right) by conditioning on NN. In particular, we have

g(𝒙)\displaystyle g\left(\boldsymbol{x}\right) =\displaystyle= n=0𝖯𝗋{N=n}𝖯𝗋{min𝑿Φ𝒟(ψ,τ)s^(𝑿)>s^(𝒙)|N=n}\displaystyle\sum_{n=0}^{\infty}\mathsf{Pr}\left\{N=n\right\}\mathsf{Pr}\left\{\min_{\boldsymbol{X}\in\Phi\cap\mathcal{D}\left(\psi,\tau\right)}\widehat{s}\left(\boldsymbol{X}\right)>\widehat{s}\left(\boldsymbol{x}\right)\big{|}N=n\right\} (99)
=\displaystyle= n=0(λπ(τ2ψ2))neλπ(τ2ψ2)n!𝖯𝗋{s^(𝑼)>s^(𝒙)}n\displaystyle\sum_{n=0}^{\infty}\frac{\left(\lambda\pi\left(\tau^{2}-\psi^{2}\right)\right)^{n}{\rm e}^{-\lambda\pi\left(\tau^{2}-\psi^{2}\right)}}{n!}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)>\widehat{s}\left(\boldsymbol{x}\right)\right\}^{n}
=\displaystyle= exp(λ2πdψ𝖯𝗋{s^(𝑼)s^(𝒙)}),\displaystyle\exp\left(-\lambda 2\pi d\psi\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)\leq\widehat{s}\left(\boldsymbol{x}\right)\right\}\right),

where 𝑼\boldsymbol{U} is uniformly distributed over 𝒟(ψ,τ)\mathcal{D}\left(\psi,\tau\right). We observe that the second case in Lemma 4 applies to calculate 𝖯𝗋{s^(𝑼)s^(𝒙)}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)\leq\widehat{s}\left(\boldsymbol{x}\right)\right\} since ψ2+d2s^(𝒙)ψ+d\sqrt{\psi^{2}+d^{2}}\leq\widehat{s}\left(\boldsymbol{x}\right)\leq\psi+d. Defining the function P(𝒙)=2πdψ𝖯𝗋{s^(𝑼)s^(𝒙)}P\left(\boldsymbol{x}\right)=2\pi d\psi\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\right)\leq\widehat{s}\left(\boldsymbol{x}\right)\right\} and using Lemma 4, we obtain

P(𝒙)=((s^(𝒙))2ψ2)(π2θ)d2sin(2θ)V(𝒙,d)+V(𝒙,dsin(θ))\displaystyle P\left(\boldsymbol{x}\right)=\left(\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}-\psi^{2}\right)\left(\pi-2\theta\right)-d^{2}\sin\left(2\theta\right)-V\left(\boldsymbol{x},d\right)+V\left(\boldsymbol{x},d\sin\left(\theta\right)\right) (100)

after some manipulations, where V(𝒙,y)=2y(s^(𝒙))2y2+2(s^(𝒙))2arctan(y(s^(𝒙))2y2)V\left(\boldsymbol{x},y\right)=2y\sqrt{\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}-y^{2}}+2\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}\arctan\left(\frac{y}{\sqrt{\left(\widehat{s}\left(\boldsymbol{x}\right)\right)^{2}-y^{2}}}\right) and θ\theta is the angle of 𝒙\boldsymbol{x} restricted to [0,π2]\left[0,\frac{\pi}{2}\right]. Let Ψmid=𝑿mid\Psi_{\rm mid}=\left\|\boldsymbol{X}_{\rm mid}\right\|, Θmid\Theta_{\rm mid} be the angle of 𝑿mid\boldsymbol{X}_{\rm mid} and fNN(ψ)f_{\rm NN}\left(\psi\right) be the pdf of Ψmid\Psi_{\rm mid}. Switching to polar coordinates, using the symmetry in the problem and writing g(𝒙)g\left(\boldsymbol{x}\right) and P(𝒙)P\left(\boldsymbol{x}\right) in polar coordinates as g(ψ,θ)g\left(\psi,\theta\right) and P(ψ,θ)P\left(\psi,\theta\right) with a slight abuse of notation, it can be seen that

𝖯𝗋{𝑿mid=𝑿opt}\displaystyle\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\right\} =\displaystyle= 12π02π0g(ψ,θ)fNN(ψ)dψdθ\displaystyle\frac{1}{2\pi}\int_{0}^{2\pi}\int_{0}^{\infty}g\left(\psi,\theta\right)f_{\rm NN}\left(\psi\right){\rm d}\psi{\rm d}\theta
=\displaystyle= 2π0π20exp(λP(ψ,θ))fNN(ψ)dψdθ\displaystyle\frac{2}{\pi}\int_{0}^{\frac{\pi}{2}}\int_{0}^{\infty}\exp\left(-\lambda P\left(\psi,\theta\right)\right)f_{\rm NN}\left(\psi\right){\rm d}\psi{\rm d}\theta
=\displaystyle= 4𝖤[exp(λP(𝑿mid))𝟣{0Θmidπ2}],\displaystyle 4\mathsf{E}\left[\exp\left(-\lambda P\left(\boldsymbol{X}_{\rm mid}\right)\right)\mathsf{1}_{\left\{0\leq\Theta_{\rm mid}\leq\frac{\pi}{2}\right\}}\right],

which concludes the proof.

Appendix D Upper Bounds on 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right)

We will only obtain the bound on 𝖱ave(𝒫opt){\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right) for the Rayleigh fading scenario since the derivation for the no-fading case is similar. In the Rayleigh fading case, we can express 𝖱Φ(𝒫){\sf R}_{\Phi}\left(\mathcal{P}\right) for any given relay selection policy 𝒫\mathcal{P} as

𝖱Φ(𝒫)\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}\right) =\displaystyle= 12𝖤[log2(1+𝖲𝖭𝖱|H|2G(s^(𝑿𝒫)))|Φ]\displaystyle\frac{1}{2}\mathsf{E}\left[\log_{2}\left(1+{\sf SNR}\left|H\right|^{2}G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)\right)\Big{|}\Phi\right] (101)
=\displaystyle= 12ln20ln(1+𝖲𝖭𝖱G(s^(𝑿𝒫))x)exdx\displaystyle\frac{1}{2\ln 2}\int_{0}^{\infty}\ln\left(1+{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)x\right){\rm e}^{-x}{\rm d}x
=(a)\displaystyle\stackrel{{\scriptstyle\rm(a)}}{{=}} 12ln20𝖲𝖭𝖱G(s^(𝑿𝒫))1+𝖲𝖭𝖱G(s^(𝑿𝒫))xexdx\displaystyle\frac{1}{2\ln 2}\int_{0}^{\infty}\frac{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)}{1+{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)x}{\rm e}^{-x}{\rm d}x
=(b)\displaystyle\stackrel{{\scriptstyle\rm(b)}}{{=}} 12ln2e1𝖲𝖭𝖱G(s^(𝑿𝒫))1et𝖲𝖭𝖱G(s^(𝑿𝒫))tdt\displaystyle\frac{1}{2\ln 2}{\rm e}^{\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)}}\int_{1}^{\infty}\frac{{\rm e}^{-\frac{t}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)}}}{t}{\rm d}t
=\displaystyle= 12ln2f(1𝖲𝖭𝖱G(s^(𝑿𝒫))),\displaystyle\frac{1}{2\ln 2}f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)}\right),

where (a) follows from integration-by-parts and (b) follows from the change of variables with t=1+𝖲𝖭𝖱G(s^(𝑿𝒫))xt=1+{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\mathcal{P}}\right)\right)x. Hence, the difference between 𝖱Φ(𝒫opt){\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right) and 𝖱Φ(𝒫mid){\sf R}_{\Phi}\left(\mathcal{P}_{\rm mid}\right) can be written as

𝖱Φ(𝒫opt)𝖱Φ(𝒫mid)=12ln2𝟣optc(f(1𝖲𝖭𝖱G(s^(𝑿opt)))f(1𝖲𝖭𝖱G(s^(𝑿mid)))),\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right)-{\sf R}_{\Phi}\left(\mathcal{P}_{\rm mid}\right)=\frac{1}{2\ln 2}\mathsf{1}_{\mathcal{E}_{\rm opt}^{c}}\left(f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm opt}\right)\right)}\right)-f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\right)}\right)\right),

where optc\mathcal{E}_{\rm opt}^{c} is the event optc{𝑿mid𝑿opt}\mathcal{E}_{\rm opt}^{c}\equiv\left\{\boldsymbol{X}_{\rm mid}\neq\boldsymbol{X}_{\rm opt}\right\}. The function f(x)=exE1(x)=ex1etxt𝑑tf(x)={\rm e}^{x}\text{E}_{1}(x)={\rm e}^{x}\int_{1}^{\infty}\frac{{\rm e}^{-tx}}{t}dt is monotone decreasing for x>0x>0 since the range of integration is [1,)\left.\left[1,\infty\right)\right. and s^(𝑿opt)Ψmid2+d2\widehat{s}\left(\boldsymbol{X}_{\rm opt}\right)\geq\sqrt{\Psi^{2}_{\rm mid}+d^{2}} by using the arguments in the proof of Theorem 1. Thus, we can upper bound the difference 𝖱Φ(𝒫opt)𝖱Φ(𝒫mid){\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right)-{\sf R}_{\Phi}\left(\mathcal{P}_{\rm mid}\right) as

𝖱Φ(𝒫opt)𝖱Φ(𝒫mid)\displaystyle{\sf R}_{\Phi}\left(\mathcal{P}_{\rm opt}\right)-{\sf R}_{\Phi}\left(\mathcal{P}_{\rm mid}\right)
12ln2𝟣optc(f(1𝖲𝖭𝖱G(Ψmid2+d2))f(1𝖲𝖭𝖱G(s^(𝑿mid)))).\displaystyle\leq\frac{1}{2\ln 2}\mathsf{1}_{\mathcal{E}_{\rm opt}^{c}}\left(f\left(\frac{1}{{\sf SNR}\cdot G\left(\sqrt{\Psi^{2}_{\rm mid}+d^{2}}\right)}\right)-f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\right)}\right)\right). (102)

We let ΔΦ=12ln2𝟣optc(f(1𝖲𝖭𝖱G(Ψmid2+d2))f(1𝖲𝖭𝖱G(s^(𝑿mid))))\Delta_{\Phi}=\frac{1}{2\ln 2}\mathsf{1}_{\mathcal{E}_{\rm opt}^{c}}\left(f\left(\frac{1}{{\sf SNR}\cdot G\left(\sqrt{\Psi^{2}_{\rm mid}+d^{2}}\right)}\right)-f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\right)}\right)\right) and Δave=𝖤[ΔΦ]\Delta_{\rm ave}=\mathsf{E}\left[\Delta_{\Phi}\right]. Then, by averaging both sides of (102), we obtain the upper bound 𝖱ave(𝒫opt)𝖱ave(𝒫mid)+Δave{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm opt}\right)\leq{\sf R}_{\rm ave}\left(\mathcal{P}_{\rm mid}\right)+\Delta_{\rm ave}. The expression for Δave\Delta_{\rm ave} is derived by first conditioning on 𝑿mid\boldsymbol{X}_{\rm mid} and then averaging over 𝑿mid\boldsymbol{X}_{\rm mid} as below:

Δave=𝖤[𝖤[ΔΦ|𝑿mid]]\displaystyle\Delta_{\rm ave}=\mathsf{E}\left[\mathsf{E}\left[\Delta_{\Phi}\big{|}\boldsymbol{X}_{\rm mid}\right]\right]
=12ln2𝖤[(f(1𝖲𝖭𝖱G(Ψmid2+d2))\displaystyle=\frac{1}{2\ln 2}\mathsf{E}\left[\left(f\left(\frac{1}{{\sf SNR}\cdot G\left(\sqrt{\Psi^{2}_{\rm mid}+d^{2}}\right)}\right)\right.\right.
f(1𝖲𝖭𝖱G(s^(𝑿mid))))𝖤[𝟣optc|𝑿mid]]\displaystyle\left.\left.-f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\right)}\right)\rule{0.0pt}{25.6073pt}\right)\mathsf{E}\left[\mathsf{1}_{\mathcal{E}_{\rm opt}^{c}}\big{|}\boldsymbol{X}_{\rm mid}\right]\right]
=2ln2𝖤[(1eλP(𝑿mid))(f(1𝖲𝖭𝖱G(Ψmid2+d2))\displaystyle=\frac{2}{\ln 2}\mathsf{E}\left[\left(1-{\rm e}^{-\lambda P\left(\boldsymbol{X}_{\rm mid}\right)}\right)\left(f\left(\frac{1}{{\sf SNR}\cdot G\left(\sqrt{\Psi_{\rm mid}^{2}+d^{2}}\right)}\right)\right.\right.
f(1𝖲𝖭𝖱G(s^(𝑿mid))))𝟣{0Θmidπ2}],\displaystyle\left.\left.-f\left(\frac{1}{{\sf SNR}\cdot G\left(\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right)\right)}\right)\rule{0.0pt}{25.6073pt}\right)\mathsf{1}_{\left\{0\leq\Theta_{\rm mid}\leq\frac{\pi}{2}\right\}}\right],

where the last equality follows from the symmetry of s^(𝑿mid)\widehat{s}\left(\boldsymbol{X}_{\rm mid}\right) over [0,2π)\left.\left[0,2\pi\right)\right. with respect to the angle of 𝑿mid\boldsymbol{X}_{\rm mid} and from the proof of Theorem 3 where it was shown that the conditional probability 𝖯𝗋{𝑿mid=𝑿opt|𝑿mid=𝒙}\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\big{|}\boldsymbol{X}_{\rm mid}=\boldsymbol{x}\right\} is equal to 𝖯𝗋{𝑿mid=𝑿opt|𝑿mid=𝒙}=eλP(𝒙)\mathsf{Pr}\left\{\boldsymbol{X}_{\rm mid}=\boldsymbol{X}_{\rm opt}\big{|}\boldsymbol{X}_{\rm mid}=\boldsymbol{x}\right\}={\rm e}^{-\lambda P\left(\boldsymbol{x}\right)} when the angle of 𝒙\boldsymbol{x} is restricted to [0,π2]\left[0,\frac{\pi}{2}\right].

Appendix E Proof of Theorem 4

We first divide the relay locations into two types:

Φright=Φright2andΦleft=Φleft2,\displaystyle\Phi_{\rm right}=\Phi\bigcap\mathbb{R}^{2}_{\rm right}\,\,\text{and}\,\,\Phi_{\rm left}=\Phi\bigcap\mathbb{R}^{2}_{\rm left}, (103)

where right2={(x1,x2)2:x10}\mathbb{R}^{2}_{\rm right}=\left\{\left(x_{1},x_{2}\right)^{\top}\in\mathbb{R}^{2}:x_{1}\geq 0\right\} and left2={(x1,x2)2:x1<0}\mathbb{R}^{2}_{\rm left}=\left\{\left(x_{1},x_{2}\right)^{\top}\in\mathbb{R}^{2}:x_{1}<0\right\}. That is, the relays in Φright\Phi_{\rm right} are closer to the destination node, whereas the ones in Φleft\Phi_{\rm left} are closer to the source node. Then, we define

Γoptrightmin𝑿Φrights^(𝑿)andΓoptleftmin𝑿Φlefts^(𝑿).\displaystyle\Gamma_{\rm opt}^{\rm right}\triangleq\min_{\boldsymbol{X}\in\Phi_{\rm right}}\widehat{s}\left(\boldsymbol{X}\right)\,\,\text{and}\,\,\Gamma_{\rm opt}^{\rm left}\triangleq\min_{\boldsymbol{X}\in\Phi_{\rm left}}\widehat{s}\left(\boldsymbol{X}\right). (104)

Due to stationarity of HPPPs and symmetry of the problem, Γoptright\Gamma_{\rm opt}^{\rm right} and Γoptleft\Gamma_{\rm opt}^{\rm left} are identically distributed random variables. Further, they are also independent due to the complete randomness property of Poisson point processes [76]. Hence, it will be enough to obtain the cdf of Γoptright\Gamma_{\rm opt}^{\rm right} to prove Theorem 4 since Γopt=min{Γoptright,Γoptleft}\Gamma_{\rm opt}=\min\left\{\Gamma_{\rm opt}^{\rm right},\Gamma_{\rm opt}^{\rm left}\right\}. More specifically, FΓopt(γ)=1(1FΓoptright(γ))2F_{\Gamma_{\rm opt}}\left(\gamma\right)=1-\left(1-F_{\Gamma_{\rm opt}^{\rm right}}\left(\gamma\right)\right)^{2}. To obtain FΓoptright(γ)F_{\Gamma_{\rm opt}^{\rm right}}\left(\gamma\right), we further define Φright,τΦright(𝟎,τ)\Phi_{{\rm right},\tau}\triangleq\Phi_{\rm right}\cap\mathcal{B}\left(\boldsymbol{0},\tau\right) and Γopt,τrightmin𝑿Φright,τs^(𝑿)\Gamma_{{\rm opt},\tau}^{\rm right}\triangleq\min_{\boldsymbol{X}\in\Phi_{{\rm right},\tau}}\widehat{s}\left(\boldsymbol{X}\right), where (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right) is the closed disc centered around the origin 𝟎\boldsymbol{0} and having radius τ\tau. We note that Γopt,τright\Gamma_{{\rm opt},\tau}^{\rm right} converges almost surely to Γoptright\Gamma_{\rm opt}^{\rm right} as τ\tau tends to infinity. Thus, the cdf of Γopt,τright\Gamma_{{\rm opt},\tau}^{\rm right} will also converge to the cdf of Γoptright\Gamma_{\rm opt}^{\rm right} pointwise as τ\tau tends to infinity [77]. We will derive the cdf of Γoptright\Gamma_{\rm opt}^{\rm right} by first obtaining the cdf of Γopt,τright\Gamma_{{\rm opt},\tau}^{\rm right} and then taking the limit τ\tau\rightarrow\infty.

Let NN be the number of relays in Φright,τ\Phi_{{\rm right},\tau}. Given the event {N=n}\left\{N=n\right\} for n1n\geq 1, all the relays in Φright,τ\Phi_{{\rm right},\tau} will be uniformly distributed over the half-disc centered at 𝟎\boldsymbol{0}, having radius τ\tau and containing only those points of 2\mathbb{R}^{2} with non-negative first coordinates. Let 𝑼\boldsymbol{U} be such a uniformly distributed random relay location. Let also Γ=s^(𝑼)\Gamma=\widehat{s}\left(\boldsymbol{U}\right), which is equal to the distance between 𝑼\boldsymbol{U} and the source node. Γ\Gamma can be written as Γ=Ψ2+2dΨcosΘ+d2\Gamma=\sqrt{\Psi^{2}+2d\Psi\cos\Theta+d^{2}} by using the law of cosines, where Ψ=𝑼\Psi=\left\|\boldsymbol{U}\right\| and Θ\Theta is the angle between the positive xx-axis and the line segment connecting 𝟎\boldsymbol{0} and 𝑼\boldsymbol{U}. Θ\Theta is uniformly distributed over [π2,π2]\left[-\frac{\pi}{2},\frac{\pi}{2}\right], and independent of Ψ\Psi due to the uniformly distributed nature of 𝑼\boldsymbol{U}. Hence, the conditional cdf of Γ\Gamma given Θ=θ\Theta=\theta can be expressed as

FΓ|Θ(γ|θ)=𝖯𝗋{Ψγ2d2sin2θdcosθ}\displaystyle F_{\Gamma|\Theta}\left(\gamma|\theta\right)=\mathsf{Pr}\left\{\Psi\leq\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right\} (105)

for dγτ2+2dτcosθ+d2d\leq\gamma\leq\sqrt{\tau^{2}+2d\tau\cos\theta+d^{2}}, where the last equation follows from the fact that the convex quadratic function f(x)=x2+2dcosθx+d2γ2f(x)=x^{2}+2d\cos\theta x+d^{2}-\gamma^{2} has only one positive root at γ2d2sin2θdcosθ\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta when γd\gamma\geq d. For γ<d\gamma<d, we have FΓ|Θ(γ|θ)=0F_{\Gamma|\Theta}\left(\gamma|\theta\right)=0 since Γ\Gamma is always greater than or equal to dd. For γ>τ2+2dτcosθ+d2\gamma>\sqrt{\tau^{2}+2d\tau\cos\theta+d^{2}}, we have FΓ|Θ(γ|θ)=1F_{\Gamma|\Theta}\left(\gamma|\theta\right)=1 since Γ\Gamma is always smaller than or equal to τ2+2dτcosθ+d2\sqrt{\tau^{2}+2d\tau\cos\theta+d^{2}} when Θ=θ\Theta=\theta. As a result, using the cdf of Ψ\Psi, which is equal to FΨ(ψ)=ψ2τ2F_{\Psi}\left(\psi\right)=\frac{\psi^{2}}{\tau^{2}}, we have

FΓ|Θ(γ|θ)={0 if γ<d(γ2d2sin2θdcosθ)2τ2 if dγτ2+2dτcosθ+d21 if γ>τ2+2dτcosθ+d2.\displaystyle F_{\Gamma|\Theta}(\gamma|\theta)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ \frac{\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)^{2}}{\tau^{2}}&\mbox{ if }d\leq\gamma\leq\sqrt{\tau^{2}+2d\tau\cos\theta+d^{2}}\\ 1&\mbox{ if }\gamma>\sqrt{\tau^{2}+2d\tau\cos\theta+d^{2}}\end{array}\right.. (109)
FΓ(γ)={0 if γ<d2πτ2(γ2arcsec(γd)dγ2d2) if dγτ2+d22γ2πτ2(arcsec(2dττ2+d2γ2)arctan(4d2τ2(τ2+d2γ2)2τ2d2+γ2))2arccsc(2dττ2+d2γ2)π(τd+γ)(τ+dγ)(dτ+γ)(τ+d+γ)πτ2 if τ2+d2<γτ+d1 if γ>τ+d.F_{\Gamma}(\gamma)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ \frac{2}{\pi\tau^{2}}\left(\gamma^{2}\operatorname{arcsec}\left(\frac{\gamma}{d}\right)-d\sqrt{\gamma^{2}-d^{2}}\right)&\mbox{ if }d\leq\gamma\leq\sqrt{\tau^{2}+d^{2}}\\ \frac{2\gamma^{2}}{\pi\tau^{2}}\left(\rule{0.0pt}{15.6491pt}\operatorname{arcsec}\left(-\frac{2d\tau}{\tau^{2}+d^{2}-\gamma^{2}}\right)\right.&\\ \hskip 28.45274pt\left.-\arctan\left(\frac{\sqrt{4d^{2}\tau^{2}-\left(\tau^{2}+d^{2}-\gamma^{2}\right)^{2}}}{\tau^{2}-d^{2}+\gamma^{2}}\right)\right)&\\ \hskip 28.45274pt-\frac{2\operatorname{arccsc}\left(\frac{2d\tau}{\tau^{2}+d^{2}-\gamma^{2}}\right)}{\pi}&\\ \hskip 28.45274pt-\frac{\sqrt{(\tau-d+\gamma)(\tau+d-\gamma)(d-\tau+\gamma)(\tau+d+\gamma)}}{\pi\tau^{2}}&\mbox{ if }\sqrt{\tau^{2}+d^{2}}<\gamma\leq\tau+d\\ 1&\mbox{ if }\gamma>\tau+d\end{array}.\right. (110)

 

We will obtain FΓ(γ)F_{\Gamma}(\gamma) by averaging (109) over Θ\Theta. To this end, we need to consider four cases separately. If γ<d\gamma<d, then FΓ|Θ(γ|θ)=0F_{\Gamma|\Theta}(\gamma|\theta)=0 for all θ[π2,π2]\theta\in\left[-\frac{\pi}{2},\frac{\pi}{2}\right]. Hence, FΓ(γ)=0F_{\Gamma}(\gamma)=0 when γ<d\gamma<d. If dγτ2+d2d\leq\gamma\leq\sqrt{\tau^{2}+d^{2}}, the second condition in (109) is always satisfied, and we have

FΓ(γ)=2πτ20π2(γ2d2sin2θdcosθ)2dθ\displaystyle F_{\Gamma}(\gamma)=\frac{2}{\pi\tau^{2}}\int_{0}^{\frac{\pi}{2}}\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)^{2}{\rm d}\theta

for this range of γ\gamma. If τ2+d2<γτ+d\sqrt{\tau^{2}+d^{2}}<\gamma\leq\tau+d, the second and third conditions in (109) are satisfied for θ[θ,θ]\theta\in\left[-\theta^{\star},\theta^{\star}\right] and θ[π2,θ)(θ,π2]\theta\in\left.\left[-\frac{\pi}{2},-\theta^{\star}\right)\right.\cup\left.\left(\theta^{\star},\frac{\pi}{2}\right]\right., respectively, where θ=arccos(γ2τ2d22dτ)\theta^{\star}=\arccos\left(\frac{\gamma^{2}-\tau^{2}-d^{2}}{2d\tau}\right). Thus, we have

FΓ(γ)=2πτ20θ(γ2d2sin2θdcosθ)2dθ+12θπ\displaystyle F_{\Gamma}(\gamma)=\frac{2}{\pi\tau^{2}}\int_{0}^{\theta^{\star}}\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)^{2}{\rm d}\theta+1-\frac{2\theta^{\star}}{\pi}

for this range of γ\gamma. Finally, if γ>τ+d\gamma>\tau+d, then FΓ|Θ(γ|θ)=1F_{\Gamma|\Theta}(\gamma|\theta)=1 for all θ[π2,π2]\theta\in\left[-\frac{\pi}{2},\frac{\pi}{2}\right], and therefore FΓ(γ)=1F_{\Gamma}(\gamma)=1 if γ>τ+d\gamma>\tau+d. Combining all four cases and evaluating the integrals, we obtain FΓ(γ)F_{\Gamma}(\gamma) as in (110). Using FΓ(γ)F_{\Gamma}(\gamma), we obtain

FΓopt,τright(γ)\displaystyle F_{\Gamma_{{\rm opt},\tau}^{\rm right}}(\gamma) =\displaystyle= n=0(1(1FΓ(γ))n)𝖯𝗋{N=n}\displaystyle\sum_{n=0}^{\infty}\left(1-\left(1-F_{\Gamma}\left(\gamma\right)\right)^{n}\right)\mathsf{Pr}\left\{N=n\right\} (111)
=\displaystyle= 1n=0(1FΓ(γ))n(λπτ22)neλπτ22n!\displaystyle 1-\sum_{n=0}^{\infty}\left(1-F_{\Gamma}\left(\gamma\right)\right)^{n}\frac{\left(\frac{\lambda\pi\tau^{2}}{2}\right)^{n}{\rm e}^{-\frac{\lambda\pi\tau^{2}}{2}}}{n!}
=\displaystyle= 1eλπτ22FΓ(γ).\displaystyle 1-{\rm e}^{-\frac{\lambda\pi\tau^{2}}{2}F_{\Gamma}\left(\gamma\right)}.

As stated earlier, FΓoptright(γ)=limτFΓopt,τright(γ)F_{\Gamma_{{\rm opt}}^{\rm right}}(\gamma)=\lim_{\tau\rightarrow\infty}F_{\Gamma_{{\rm opt},\tau}^{\rm right}}(\gamma) since Γopt,τright\Gamma_{{\rm opt},\tau}^{\rm right} converges to Γoptright\Gamma_{{\rm opt}}^{\rm right} almost surely. Thus, by rearranging the terms in (110) and using (111), we have

FΓoptright(γ)\displaystyle F_{\Gamma_{{\rm opt}}^{\rm right}}(\gamma) =\displaystyle= limτFΓopt,τright(γ)\displaystyle\lim_{\tau\rightarrow\infty}F_{\Gamma_{{\rm opt},\tau}^{\rm right}}(\gamma) (114)
=\displaystyle= 1limτeλπτ22FΓ(γ)\displaystyle 1-\lim_{\tau\rightarrow\infty}{\rm e}^{-\frac{\lambda\pi\tau^{2}}{2}F_{\Gamma}\left(\gamma\right)}
=\displaystyle= {0 if γ<d1eλd2((γd)2arcsec(γd)(γd)21) if γd.\displaystyle\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ 1-{\rm e}^{-\lambda d^{2}\left(\left(\frac{\gamma}{d}\right)^{2}\operatorname{arcsec}\left(\frac{\gamma}{d}\right)-\sqrt{\left(\frac{\gamma}{d}\right)^{2}-1}\right)}&\mbox{ if }\gamma\geq d\end{array}.\right.

Finally, using (114) and the identity FΓopt(γ)=1(1FΓoptright(γ))2F_{\Gamma_{\rm opt}}\left(\gamma\right)=1-\left(1-F_{\Gamma_{\rm opt}^{\rm right}}\left(\gamma\right)\right)^{2}, we conclude the proof.

Appendix F Proof of Theorem 5

To prove this theorem, we first focus on the relay nodes located inside the closed disc (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right). Let μ(T,τ)\mu\left(T,\tau\right) be the average number of relays located in (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right) and feeding their channel quality indicators back to the source node. μ(T,τ)\mu\left(T,\tau\right) is given by

μ(T,τ)=𝖤[𝑿Φ(𝟎,τ)𝟣{s^(𝑿)T}].\displaystyle\mu\left(T,\tau\right)=\mathsf{E}\left[\sum_{\boldsymbol{X}\in\Phi\cap\mathcal{B}\left(\boldsymbol{0},\tau\right)}\mathsf{1}_{\left\{\widehat{s}\left(\boldsymbol{X}\right)\leq T\right\}}\right].

Using the monotone convergence theorem, it can be seen that μ(T)=limτμ(T,τ)\mu\left(T\right)=\lim_{\tau\rightarrow\infty}\mu\left(T,\tau\right). Let NN be the number of relays in Φ(𝟎,τ)\Phi\cap\mathcal{B}\left(\boldsymbol{0},\tau\right). Given the event {N=n}\left\{N=n\right\}, all the relays are independently and uniformly distributed over (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right), and hence 𝖤[𝑿Φ(𝟎,τ)𝟣{s^(𝑿)T}|N=n]=n𝖯𝗋{s^(𝑼T)}\mathsf{E}\left[\sum_{\boldsymbol{X}\in\Phi\cap\mathcal{B}\left(\boldsymbol{0},\tau\right)}\mathsf{1}_{\left\{\widehat{s}\left(\boldsymbol{X}\right)\leq T\right\}}\Big{|}N=n\right]=n\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\leq T\right)\right\}, where 𝑼\boldsymbol{U} is a generic random variable uniformly distributed over (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right). Using this observation, we can write μ(T,τ)\mu\left(T,\tau\right) as

μ(T,τ)=λπτ2𝖯𝗋{s^(𝑼T)}.\displaystyle\mu\left(T,\tau\right)=\lambda\pi\tau^{2}\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\leq T\right)\right\}. (115)

We will use Lemma 4 to conclude the proof, and it is enough to focus only on the case where dTd2+τ2d\leq T\leq\sqrt{d^{2}+\tau^{2}}. In particular, it can be seen by using this lemma that 𝖯𝗋{s^(𝑼T)}=0\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\leq T\right)\right\}=0 for all values of TT smaller than dd. Therefore, μ(T)=limτμ(T,τ)=0\mu\left(T\right)=\lim_{\tau\rightarrow\infty}\mu\left(T,\tau\right)=0 for T<dT<d. For other cases of this lemma where Td2+τ2T\geq\sqrt{d^{2}+\tau^{2}}, the threshold value grows without any bound when τ\tau tends to infinity, which is equivalent to the all-feedback case investigated in Section VI.

For dTd2+τ2d\leq T\leq\sqrt{d^{2}+\tau^{2}}, we have

𝖯𝗋{s^(𝑼T)}=T2τ22dT2d2πτ22T2πτ2arctan(dT2d2)\displaystyle\mathsf{Pr}\left\{\widehat{s}\left(\boldsymbol{U}\leq T\right)\right\}=\frac{T^{2}}{\tau^{2}}-\frac{2d\sqrt{T^{2}-d^{2}}}{\pi\tau^{2}}-\frac{2T^{2}}{\pi\tau^{2}}\arctan\left(\frac{d}{\sqrt{T^{2}-d^{2}}}\right)

by using Lemma 4. As a result,

μ(T,τ)=λπT22dλT2d22T2λarctan(dT2d2).\displaystyle\mu\left(T,\tau\right)=\lambda\pi T^{2}-2d\lambda\sqrt{T^{2}-d^{2}}-2T^{2}\lambda\arctan\left(\frac{d}{\sqrt{T^{2}-d^{2}}}\right).

Taking the limit as τ\tau tends to infinity, we obtain (41). To obtain the distribution of NFBN_{\rm FB}, we first observe that the sum 𝑿Φ(𝟎,τ)𝟣{s^(𝑿)T}\sum_{\boldsymbol{X}\in\Phi\cap\mathcal{B}\left(\boldsymbol{0},\tau\right)}\mathsf{1}_{\left\{\widehat{s}\left(\boldsymbol{X}\right)\leq T\right\}} has the characteristic function φτ(t)=exp(μ(τ,T)(eȷt1))\varphi_{\tau}(t)=\exp\left(\mu\left(\tau,T\right)\left({\rm e}^{\jmath t}-1\right)\right), where ȷ=1\jmath=\sqrt{-1}. Since NFB=limτ𝑿Φ(𝟎,τ)𝟣{s^(𝑿)T}N_{\rm FB}=\lim_{\tau\rightarrow\infty}\sum_{\boldsymbol{X}\in\Phi\cap\mathcal{B}\left(\boldsymbol{0},\tau\right)}\mathsf{1}_{\left\{\widehat{s}\left(\boldsymbol{X}\right)\leq T\right\}} almost surely, we conclude that NFBN_{\rm FB} has a Poisson distribution with mean μ(T)\mu\left(T\right) [77].

Appendix G Derivation of Distribution Functions for Γmid\Gamma_{\rm mid} and ΓC2\Gamma_{\rm C2\star} with {D,S}\star\in\left\{\rm D,S\right\}

In this appendix, we will derive distribution functions for Γmid\Gamma_{\rm mid} and ΓC2\Gamma_{\rm C2\star}, {D,S}\star\in\left\{\rm D,S\right\}, given in (59) - (62). We will start with FΓmid(γ)F_{\Gamma_{\rm mid}}\left(\gamma\right) and fΓmid(γ)f_{\Gamma_{\rm mid}}\left(\gamma\right).

G-A Probability Distribution for Γmid\Gamma_{\rm mid}

Let 𝑿mid\boldsymbol{X}_{\rm mid} be the closest point of Φ\Phi to the origin, and Ψmid=𝑿mid\Psi_{\rm mid}=\left\|\boldsymbol{X}_{\rm mid}\right\| and Θmid\Theta_{\rm mid} be the angle of 𝑿mid\boldsymbol{X}_{\rm mid}. Ψmid\Psi_{\rm mid} and Θmid\Theta_{\rm mid} are independent of each other, with Ψmid\Psi_{\rm mid} having the nearest-neighbor pdf fNN(ψ)f_{\rm NN}\left(\psi\right) for HPPPs as its pdf, and Θmid\Theta_{\rm mid} being uniformly distributed over [0,2π)[0,2\pi). Let \mathcal{E} be the event ={Θ[0,π2]}\mathcal{E}=\left\{\Theta\in\left[0,\frac{\pi}{2}\right]\right\}. Due to symmetry in the problem, we have FΓmid(γ)=FΓmid|(γ)F_{\Gamma_{\rm mid}}\left(\gamma\right)=F_{\Gamma_{\rm mid}|\mathcal{E}}\left(\gamma\right). Hence, we will focus on FΓmid|(γ)F_{\Gamma_{\rm mid}|\mathcal{E}}\left(\gamma\right) for the rest of the derivation below.

We first observe that Γmid\Gamma_{\rm mid} is always greater than dd, and hence FΓmid|(γ)=0F_{\Gamma_{\rm mid}|\mathcal{E}}\left(\gamma\right)=0 for γ<d\gamma<d. For γd\gamma\geq d, we define the function g(γ,ψ)g\left(\gamma,\psi\right) as

g(γ,ψ)=FΓmid|,Ψmid(γ|ψ).\displaystyle g\left(\gamma,\psi\right)=F_{\Gamma_{\rm mid}|\mathcal{E},\Psi_{\rm mid}}\left(\gamma|\psi\right).

Given Ψmid=ψ\Psi_{\rm mid}=\psi, the smallest value Γmid\Gamma_{\rm mid} can take is ψ2+d2\sqrt{\psi^{2}+d^{2}}, which is attained when Θmid=π2\Theta_{\rm mid}=\frac{\pi}{2}. The largest value, on the other hand, is ψ+d\psi+d, which is attained at Θmid=0\Theta_{\rm mid}=0. Hence, we can write g(γ,ψ)=0g\left(\gamma,\psi\right)=0 if γ<ψ2+d2\gamma<\sqrt{\psi^{2}+d^{2}}, and g(γ,ψ)=1g\left(\gamma,\psi\right)=1 if γ>ψ+d\gamma>\psi+d. For ψ2+d2γψ+d\sqrt{\psi^{2}+d^{2}}\leq\gamma\leq\psi+d, we have

g(γ,ψ)\displaystyle g\left(\gamma,\psi\right) =\displaystyle= 𝖯𝗋{ψ2+2dψcos(Θ)+d2γ2|}\displaystyle\mathsf{Pr}\left\{\psi^{2}+2d\psi\cos\left(\Theta\right)+d^{2}\leq\gamma^{2}\ \big{|}\ \mathcal{E}\right\}
=\displaystyle= 𝖯𝗋{cos(Θ)γ2ψ2d22dψ|}\displaystyle\mathsf{Pr}\left\{\cos\left(\Theta\right)\leq\frac{\gamma^{2}-\psi^{2}-d^{2}}{2d\psi}\ \Big{|}\ \mathcal{E}\right\}
=\displaystyle= 12πarccos(γ2ψ2d22dψ).\displaystyle 1-\frac{2}{\pi}\arccos\left(\frac{\gamma^{2}-\psi^{2}-d^{2}}{2d\psi}\right).

Using independence between Ψmid\Psi_{\rm mid} and Θmid\Theta_{\rm mid}, and integrating g(γ,ψ)g\left(\gamma,\psi\right) over ψ\psi with respect to fNN(ψ)f_{\rm NN}\left(\psi\right), we obtain FΓmid(γ)F_{\Gamma_{\rm mid}}\left(\gamma\right) for γd\gamma\geq d according to

FΓmid(γ)\displaystyle F_{\Gamma_{\rm mid}}\left(\gamma\right) =\displaystyle= 0g(γ,ψ)fNN(ψ)dψ\displaystyle\int_{0}^{\infty}g\left(\gamma,\psi\right)f_{\rm NN}\left(\psi\right){\rm d}\psi (116)
=\displaystyle= 0γdfNN(ψ)𝑑ψ+γdγ2d2fNN(ψ)(12πarccos(γ2ψ2d22dψ))dψ\displaystyle\int_{0}^{\gamma-d}f_{\rm NN}\left(\psi\right)d\psi+\int_{\gamma-d}^{\sqrt{\gamma^{2}-d^{2}}}f_{\rm NN}\left(\psi\right)\left(1-\frac{2}{\pi}\arccos\left(\frac{\gamma^{2}-\psi^{2}-d^{2}}{2d\psi}\right)\right){\rm d}\psi
=\displaystyle= FNN(γ2d2)2πγdγ2d2fNN(ψ)arccos(γ2ψ2d22dψ)dψ\displaystyle F_{\rm NN}\left(\sqrt{\gamma^{2}-d^{2}}\right)-\frac{2}{\pi}\int_{\gamma-d}^{\sqrt{\gamma^{2}-d^{2}}}f_{\rm NN}\left(\psi\right)\arccos\left(\frac{\gamma^{2}-\psi^{2}-d^{2}}{2d\psi}\right){\rm d}\psi
=\displaystyle= FNN(γ2d2)2λγdγ2d22ψeλπψ2arccos(γ2ψ2d22dψ)dψ\displaystyle F_{\rm NN}\left(\sqrt{\gamma^{2}-d^{2}}\right)-2\lambda\int_{\gamma-d}^{\sqrt{\gamma^{2}-d^{2}}}2\psi{\rm e}^{-\lambda\pi\psi^{2}}\arccos\left(\frac{\gamma^{2}-\psi^{2}-d^{2}}{2d\psi}\right){\rm d}\psi

Using change of variable x=ψ2x=\psi^{2} in (116) and the fact that FΓmid(γ)=0F_{\Gamma_{\rm mid}}\left(\gamma\right)=0 for γ<d\gamma<d, we arrive at

FΓmid(γ)=(FNN(γ2d2)2λ(γd)2γ2d2eλπxarccos(γ2xd22dx)dx)𝟣{γd}.\displaystyle F_{\Gamma_{\rm mid}}\left(\gamma\right)=\left(F_{\rm NN}\left(\sqrt{\gamma^{2}-d^{2}}\right)-2\lambda\int_{\left(\gamma-d\right)^{2}}^{\gamma^{2}-d^{2}}{\rm e}^{-\lambda\pi x}\arccos\left(\frac{\gamma^{2}-x-d^{2}}{2d\sqrt{x}}\right){\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}}. (117)

Using FΓmid(γ)F_{\Gamma_{\rm mid}}\left(\gamma\right) and Leibniz rule for differentiation under integral sign, we obtain

fΓmid(γ)=(4λγ(γd)2γ2d2eλπx4d2x(γ2xd2)2dx)𝟣{γd}.\displaystyle f_{\Gamma_{\rm mid}}\left(\gamma\right)=\left(4\lambda\gamma\int_{\left(\gamma-d\right)^{2}}^{\gamma^{2}-d^{2}}\frac{{\rm e}^{-\lambda\pi x}}{\sqrt{4d^{2}x-\left(\gamma^{2}-x-d^{2}\right)^{2}}}{\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}}. (118)

We note that fΓmid(γ)f_{\Gamma_{\rm mid}}\left(\gamma\right) is well defined at γ=d\gamma=d since the left and right derivatives are equal to zero.

G-B Probability Distribution for ΓC2\Gamma_{\rm C2\star} with {D,S}\star\in\left\{\rm D,S\right\}

In this part, we will derive the distribution for CQI for the closest-to-destination 𝒫C2D\mathcal{P}_{\rm C2D} and closest-to-source 𝒫C2S\mathcal{P}_{\rm C2S} relay selection policies. Due to symmetry in the problem, it is enough to focus on only one of these policies. Below, we will provide the derivation for 𝒫C2D\mathcal{P}_{\rm C2D}. Let 𝑿C2D\boldsymbol{X}_{\rm C2D} be the closest point of Φ\Phi to 𝒙d\boldsymbol{x}_{\rm d}, and ΨC2D=𝑿C2D𝒙d\Psi_{\rm C2D}=\left\|\boldsymbol{X}_{\rm C2D}-\boldsymbol{x}_{\rm d}\right\| and ΘC2D\Theta_{\rm C2D} be the angle between the positive xx-axis and the line segment connecting 𝒙d\boldsymbol{x}_{\rm d} with 𝑿C2D\boldsymbol{X}_{\rm C2D}. Let 𝒜={ΘC2D(π2,3π2)}\mathcal{A}=\left\{\Theta_{\rm C2D}\in\left(\frac{\pi}{2},\frac{3\pi}{2}\right)\right\} and ={ΘC2D[0,π2][3π2,2π)}\mathcal{E}=\left\{\Theta_{\rm C2D}\in\left[0,\frac{\pi}{2}\right]\cup\left[\frac{3\pi}{2},2\pi\right)\right\}. Using law of total probability, we can write

FΓC2D(γ)=12FΓC2D|𝒜(γ)+12FΓC2D|(γ).\displaystyle F_{\Gamma_{\rm C2D}}\left(\gamma\right)=\frac{1}{2}F_{\Gamma_{\rm C2D}|\mathcal{A}}\left(\gamma\right)+\frac{1}{2}F_{\Gamma_{\rm C2D}|\mathcal{E}}\left(\gamma\right). (119)

The derivation of FΓC2D|(γ)F_{\Gamma_{\rm C2D}|\mathcal{E}}\left(\gamma\right) follows from the above arguments we used to obtain FΓmid(γ)F_{\Gamma_{\rm mid}}\left(\gamma\right) since the geometry of the problem is exactly the same except with the change of dd with 2d2d in this case. Hence, we have

FΓC2D|(γ)=(FNN(γ24d2)2λ(γ2d)2γ24d2eλπxarccos(γ2x4d24dx)dx)𝟣{γ2d}.\displaystyle F_{\Gamma_{\rm C2D}|\mathcal{E}}\left(\gamma\right)=\left(F_{\rm NN}\left(\sqrt{\gamma^{2}-4d^{2}}\right)-2\lambda\int_{\left(\gamma-2d\right)^{2}}^{\gamma^{2}-4d^{2}}{\rm e}^{-\lambda\pi x}\arccos\left(\frac{\gamma^{2}-x-4d^{2}}{4d\sqrt{x}}\right){\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq 2d\right\}}. (120)

For derivation of FΓC2D|𝒜(γ)F_{\Gamma_{\rm C2D}|\mathcal{A}}\left(\gamma\right), we define the function g(γ,ψ)g\left(\gamma,\psi\right) as

g(γ,ψ)=FΓC2D|𝒜ΨC2D(γ|ψ).\displaystyle g\left(\gamma,\psi\right)=F_{\Gamma_{\rm C2D}|\mathcal{A}\ \Psi_{\rm C2D}}\left(\gamma|\psi\right). (121)

We will analyze the cases ψ[0,d]\psi\in[0,d] and ψ(d,)\psi\in(d,\infty) separately to obtain a functional form for g(γ,ψ)g\left(\gamma,\psi\right). We will first consider the case ψ[0,d]\psi\in[0,d]. Given ΨC2D=ψ[0,d]\Psi_{\rm C2D}=\psi\in[0,d], we have ΓC2D=ψ24dψ|cos(ΘC2D)|+4d2\Gamma_{\rm C2D}=\sqrt{\psi^{2}-4d\psi\left|\cos\left(\Theta_{\rm C2D}\right)\right|+4d^{2}}, which lies in [2dψ,ψ2+4d2)\left[2d-\psi,\sqrt{\psi^{2}+4d^{2}}\right) when ΘC2D(π2,3π2)\Theta_{\rm C2D}\in\left(\frac{\pi}{2},\frac{3\pi}{2}\right). Hence, we can write g(γ,ψ)g\left(\gamma,\psi\right) as

g(γ,ψ)={0 if γ<d2πarccos(ψ2+4d2γ24dψ)𝟣{2dγψd} if dγ<2d𝟣{0ψγ24d2}+2πarccos(ψ2+4d2γ24dψ)𝟣{γ24d2<ψd} if 2dγ<5d𝟣{ψ[0,d]} if γ5d\displaystyle g\left(\gamma,\psi\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ \frac{2}{\pi}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)\mathsf{1}_{\left\{2d-\gamma\leq\psi\leq d\right\}}&\mbox{ if }d\leq\gamma<2d\\ \mathsf{1}_{\left\{0\leq\psi\leq\sqrt{\gamma^{2}-4d^{2}}\right\}}+\frac{2}{\pi}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)\mathsf{1}_{\left\{\sqrt{\gamma^{2}-4d^{2}}<\psi\leq d\right\}}&\mbox{ if }2d\leq\gamma<\sqrt{5}d\\ \mathsf{1}_{\left\{\psi\in\left[0,d\right]\right\}}&\mbox{ if }\gamma\geq\sqrt{5}d\end{array}\right. (126)

for four different ranges of γ\gamma in this case.

Now, we consider the case ΨC2D=ψ(d,)\Psi_{\rm C2D}=\psi\in\left(d,\infty\right). In this case, ΓC2D=ψ\Gamma_{\rm C2D}=\psi when ΘC2D[πarccos(dψ),π+arccos(dψ)]\Theta_{\rm C2D}\in\left[\pi-\arccos\left(\frac{d}{\psi}\right),\pi+\arccos\left(\frac{d}{\psi}\right)\right]. On the other hand, ΓC2D=ψ24dψ|cos(ΘC2D)|+4d2\Gamma_{\rm C2D}=\sqrt{\psi^{2}-4d\psi\left|\cos\left(\Theta_{\rm C2D}\right)\right|+4d^{2}} when ΘC2D(π2,πarccos(dψ))(π+arccos(dψ),3π2)\Theta_{\rm C2D}\in\left(\frac{\pi}{2},\pi-\arccos\left(\frac{d}{\psi}\right)\right)\bigcup\left(\pi+\arccos\left(\frac{d}{\psi}\right),\frac{3\pi}{2}\right). Hence, we can write g(γ,ψ)g\left(\gamma,\psi\right) as

g(γ,ψ)={0 if γ<d2πarccos(ψ2+4d2γ24dψ)𝟣{d<ψγ} if dγ<5d𝟣{d<ψγ24d2}+2πarccos(ψ2+4d2γ24dψ)𝟣{γ24d2<ψγ} if γ>5d\displaystyle g\left(\gamma,\psi\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ \frac{2}{\pi}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)\mathsf{1}_{\left\{d<\psi\leq\gamma\right\}}&\mbox{ if }d\leq\gamma<\sqrt{5}d\\ \mathsf{1}_{\left\{d<\psi\leq\sqrt{\gamma^{2}-4d^{2}}\right\}}+\frac{2}{\pi}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)\mathsf{1}_{\left\{\sqrt{\gamma^{2}-4d^{2}}<\psi\leq\gamma\right\}}&\mbox{ if }\gamma>\sqrt{5}d\end{array}\right. (130)

for three different ranges of γ\gamma. Combining (126) and (130), we arrive at

g(γ,ψ)={0 if γ<d2πarccos(ψ2+4d2γ24dψ)𝟣{2dγ<ψγ} if dγ<2d𝟣{0ψγ24d2}+2πarccos(ψ2+4d2γ24dψ)𝟣{γ24d2<ψγ} if γ2d.\displaystyle g\left(\gamma,\psi\right)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ \frac{2}{\pi}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)\mathsf{1}_{\left\{2d-\gamma<\psi\leq\gamma\right\}}&\mbox{ if }d\leq\gamma<2d\\ \mathsf{1}_{\left\{0\leq\psi\leq\sqrt{\gamma^{2}-4d^{2}}\right\}}+\frac{2}{\pi}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)\mathsf{1}_{\left\{\sqrt{\gamma^{2}-4d^{2}}<\psi\leq\gamma\right\}}&\mbox{ if }\gamma\geq 2d\end{array}.\right. (134)

The identity in (134) gives us the complete functional representation for the function g(γ,ψ)g\left(\gamma,\psi\right). Using (134), we can write FΓC2D|𝒜(γ)F_{\Gamma_{\rm C2D}|\mathcal{A}}\left(\gamma\right) as

FΓC2D|𝒜(γ)\displaystyle F_{\Gamma_{\rm C2D}|\mathcal{A}}\left(\gamma\right)
=0g(γ,ψ)fNN(ψ)dψ\displaystyle=\int_{0}^{\infty}g\left(\gamma,\psi\right)f_{\rm NN}\left(\psi\right){\rm d}\psi
={0 if γ<d2π2dγγarccos(ψ2+4d2γ24dψ)fNN(ψ)dψ if dγ<2dFNN(γ24d2)+2πγ24d2γarccos(ψ2+4d2γ24dψ)fNN(ψ)dψ if γ2d.\displaystyle=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ \frac{2}{\pi}\int_{2d-\gamma}^{\gamma}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)f_{\rm NN}\left(\psi\right){\rm d}\psi&\mbox{ if }d\leq\gamma<2d\\ F_{\rm NN}\left(\sqrt{\gamma^{2}-4d^{2}}\right)+\frac{2}{\pi}\int_{\sqrt{\gamma^{2}-4d^{2}}}^{\gamma}\arccos\left(\frac{\psi^{2}+4d^{2}-\gamma^{2}}{4d\psi}\right)f_{\rm NN}\left(\psi\right){\rm d}\psi&\mbox{ if }\gamma\geq 2d\end{array}.\right. (138)

Using the identities arccos(x)=πarccosx\arccos\left(-x\right)=\pi-\arccos{x}, FNN(x)=0F_{\rm NN}\left(x\right)=0 for x<0x<0, and (119), (120) and (138), we arrive at

FΓC2D(γ)=(FNN(γ2d)+λ(γ2d)2γ2eλπxarccos(x+4d2γ24dx)dx)𝟣{γd}.\displaystyle F_{\Gamma_{\rm C2D}}\left(\gamma\right)=\left(F_{\rm NN}\left(\gamma-2d\right)+\lambda\int_{\left(\gamma-2d\right)^{2}}^{\gamma^{2}}{\rm e}^{-\lambda\pi x}\arccos\left(\frac{x+4d^{2}-\gamma^{2}}{4d\sqrt{x}}\right){\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}}. (139)

Using FΓC2D(γ)F_{\Gamma_{\rm C2D}}\left(\gamma\right) and Leibniz rule for differentiation under integral sign, we obtain

fΓC2D(γ)=2λγ(arccos(dγ)eλπγ2+(γ2d)2γ2eλπx16d2x(x+4d2γ2)2dx)𝟣{γd}.\displaystyle f_{\Gamma_{\rm C2D}}\left(\gamma\right)=2\lambda\gamma\left(\arccos\left(\frac{d}{\gamma}\right){\rm e}^{-\lambda\pi\gamma^{2}}+\int_{\left(\gamma-2d\right)^{2}}^{\gamma^{2}}\frac{{\rm e}^{-\lambda\pi x}}{\sqrt{16d^{2}x-\left(x+4d^{2}-\gamma^{2}\right)^{2}}}{\rm d}x\right)\mathsf{1}_{\left\{\gamma\geq d\right\}}. (140)

As for Γmid\Gamma_{\rm mid}, fΓC2D(γ)f_{\Gamma_{\rm C2D}}\left(\gamma\right) is well defined at γ=d\gamma=d since the left and right derivatives are equal to zero.

Appendix H Proof of Theorem 8

The proof of this theorem follows from the proof of Theorem 4 until (105), where we obtained the conditional cdf of Γ=s^(𝑼)\Gamma=\widehat{s}\left(\boldsymbol{U}\right) given the angle Θ\Theta of any random relay location 𝑼\boldsymbol{U} in (𝟎,τ)right2\mathcal{B}\left(\boldsymbol{0},\tau\right)\bigcap\mathbb{R}^{2}_{\rm right} and given that there are nn, n1n\geq 1, relays in Φright,τ\Phi_{{\rm right},\tau}. The following lemma establishes the independence property for the angle and magnitude of 𝑼\boldsymbol{U} to proceed with the rest of the proof for isotropic PPPs.

Lemma 5

Let NN be the number of relays in Φright,τ\Phi_{{\rm right},\tau} and 𝐔\boldsymbol{U} be any random relay location in Φright,τ\Phi_{{\rm right},\tau} given {N=n}\left\{N=n\right\} for n1n\geq 1. Then, its angle Θ\Theta is uniformly distributed over [π2,π2]\left[-\frac{\pi}{2},\frac{\pi}{2}\right] and is independent from its magnitude Ψ=𝐔\Psi=\left\|\boldsymbol{U}\right\|.

Proof:

Using the Poisson property and isotropy [76], we can express 𝖯𝗋{𝑼𝒮}\mathsf{Pr}\left\{\boldsymbol{U}\in\mathcal{S}\right\} as

𝖯𝗋{𝑼𝒮}=2Λ(𝒮)Λ((𝟎,τ))\displaystyle\mathsf{Pr}\left\{\boldsymbol{U}\in\mathcal{S}\right\}=\frac{2\Lambda\left(\mathcal{S}\right)}{\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)}

for all Borel subsets 𝒮\mathcal{S} of (𝟎,τ)right2\mathcal{B}\left(\boldsymbol{0},\tau\right)\bigcap\mathbb{R}^{2}_{\rm right}. We now consider an auxiliary random variable 𝑼~\widetilde{\boldsymbol{U}} with distribution given according to

𝖯𝗋{𝑼~𝒮}=Λ(𝒮)Λ((𝟎,τ))\displaystyle\mathsf{Pr}\left\{\widetilde{\boldsymbol{U}}\in\mathcal{S}\right\}=\frac{\Lambda\left(\mathcal{S}\right)}{\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)}

for all Borel subsets 𝒮\mathcal{S} of (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right). 𝑼~\widetilde{\boldsymbol{U}} is a spherically symmetric random variable because

𝖯𝗋{Π(𝑼~)𝒮}\displaystyle\mathsf{Pr}\left\{\Pi\left(\widetilde{\boldsymbol{U}}\right)\in\mathcal{S}\right\} =\displaystyle= 𝖯𝗋{𝑼~Π1(𝒮)}\displaystyle\mathsf{Pr}\left\{\widetilde{\boldsymbol{U}}\in\Pi^{-1}\left(\mathcal{S}\right)\right\} (141)
=\displaystyle= Λ(Π1(𝒮))Λ((𝟎,τ))\displaystyle\frac{\Lambda\left(\Pi^{-1}\left(\mathcal{S}\right)\right)}{\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)}
=\displaystyle= Λ(𝒮)Λ((𝟎,τ))\displaystyle\frac{\Lambda\left(\mathcal{S}\right)}{\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)}

for all rotations Π\Pi around the origin, where the last equality follows from the isotropy property. Hence, the magnitude Ψ~\widetilde{\Psi} of 𝑼~\widetilde{\boldsymbol{U}} is independent of its angle Θ~\widetilde{\Theta}, which is uniformly distributed over [0,2π)\left.\left[0,2\pi\right)\right., i.e., see [97, Theorem 2.3]. Now, we consider conditional distribution of 𝑼~\widetilde{\boldsymbol{U}} given Θ~[π2,π2]\widetilde{\Theta}\in\left[-\frac{\pi}{2},\frac{\pi}{2}\right]. For any Borel subset 𝒮\mathcal{S} of (𝟎,τ)right2\mathcal{B}\left(\boldsymbol{0},\tau\right)\bigcap\mathbb{R}^{2}_{\rm right}, it is equal to

𝖯𝗋{𝑼~𝒮|Θ~[π2,π2]}\displaystyle\mathsf{Pr}\left\{\widetilde{\boldsymbol{U}}\in\mathcal{S}\Big{|}\widetilde{\Theta}\in\left[-\frac{\pi}{2},\frac{\pi}{2}\right]\right\} =\displaystyle= 𝖯𝗋({𝑼~𝒮}{Θ~[π2,π2]})𝖯𝗋{Θ~[π2,π2]}\displaystyle\frac{\mathsf{Pr}\left(\left\{\widetilde{\boldsymbol{U}}\in\mathcal{S}\right\}\bigcap\left\{\widetilde{\Theta}\in\left[-\frac{\pi}{2},\frac{\pi}{2}\right]\right\}\right)}{\mathsf{Pr}\left\{\widetilde{\Theta}\in\left[-\frac{\pi}{2},\frac{\pi}{2}\right]\right\}} (142)
=\displaystyle= 2𝖯𝗋{𝑼~𝒮}\displaystyle 2\mathsf{Pr}\left\{\widetilde{\boldsymbol{U}}\in\mathcal{S}\right\}
=\displaystyle= 2Λ(𝒮)Λ((𝟎,τ)),\displaystyle\frac{2\Lambda\left(\mathcal{S}\right)}{\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)},

which is the same distribution with that of 𝑼\boldsymbol{U}. Using this identity and taking 𝒮\mathcal{S} to be 𝒮1={(ψcosθ,ψsinθ):θ[θ1,θ2],ψ[0,τ]}\mathcal{S}_{1}=\left\{\left(\psi\cos\theta,\psi\sin\theta\right):\theta\in\left[\theta_{1},\theta_{2}\right],\psi\in\left[0,\tau\right]\right\}, 𝒮2={(ψcosθ,ψsinθ):θ[π2,π2],ψ[ψ1,ψ2]}\mathcal{S}_{2}=\left\{\left(\psi\cos\theta,\psi\sin\theta\right):\theta\in\left[-\frac{\pi}{2},\frac{\pi}{2}\right],\psi\in\left[\psi_{1},\psi_{2}\right]\right\} and 𝒮3={(ψcosθ,ψsinθ):θ[θ1,θ2],ψ[ψ1,ψ2]}\mathcal{S}_{3}=\left\{\left(\psi\cos\theta,\psi\sin\theta\right):\theta\in\left[\theta_{1},\theta_{2}\right],\psi\in\left[\psi_{1},\psi_{2}\right]\right\} for π2θ1θ2π2-\frac{\pi}{2}\leq\theta_{1}\leq\theta_{2}\leq\frac{\pi}{2} and 0ψ1ψ2τ0\leq\psi_{1}\leq\psi_{2}\leq\tau, it can be seen that Ψ\Psi and Θ\Theta are independent and Θ\Theta is uniformly distributed over [π2,π2]\left[-\frac{\pi}{2},\frac{\pi}{2}\right]. ∎

Using Lemma 5, we can express FΓ|Θ(γ|θ)F_{\Gamma|\Theta}(\gamma|\theta) for isotropic PPPs according to

FΓ|Θ(γ|θ)={0 if γ<dΛ((𝟎,γ2d2sin2θdcosθ))Λ((𝟎,τ)) if dγτ2+2dτcosθ+d21 if γ>τ2+2dτcosθ+d2.\displaystyle F_{\Gamma|\Theta}(\gamma|\theta)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ \frac{\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)\right)}{\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)}&\mbox{ if }d\leq\gamma\leq\sqrt{\tau^{2}+2d\tau\cos\theta+d^{2}}\\ 1&\mbox{ if }\gamma>\sqrt{\tau^{2}+2d\tau\cos\theta+d^{2}}\end{array}\right.. (146)

We will obtain FΓ(γ)F_{\Gamma}\left(\gamma\right) by averaging (146) over the distribution of Θ\Theta by considering four different cases. Two of them are trivial. For γ<d\gamma<d, FΓ(γ)=0F_{\Gamma}\left(\gamma\right)=0, and FΓ(γ)=1F_{\Gamma}\left(\gamma\right)=1 for γ>τ+d\gamma>\tau+d. For dγτ2+d2d\leq\gamma\leq\sqrt{\tau^{2}+d^{2}}, we have

FΓ(γ)=2πΛ((𝟎,τ))0π2Λ((𝟎,γ2d2sin2θdcosθ))dθ,\displaystyle F_{\Gamma}\left(\gamma\right)=\frac{2}{\pi\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)}\int_{0}^{\frac{\pi}{2}}\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)\right){\rm d}\theta,

while we have

FΓ(γ)=1+2πΛ((𝟎,τ))0θΛ((𝟎,γ2d2sin2θdcosθ))dθ2θπ\displaystyle F_{\Gamma}\left(\gamma\right)=1+\frac{2}{\pi\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\tau\right)\right)}\int_{0}^{\theta^{\star}}\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)\right){\rm d}\theta-\frac{2\theta^{\star}}{\pi}

for τ2+d2<γτ+d\sqrt{\tau^{2}+d^{2}}<\gamma\leq\tau+d, where θ=arccos(γ2τ2d22dτ)\theta^{\star}=\arccos\left(\frac{\gamma^{2}-\tau^{2}-d^{2}}{2d\tau}\right). Using the same definitions in Appendix E, averaging over NN and taking the limit as τ\tau goes to infinity, we obtain

FΓoptright(γ)={0γ<d1e1π0π2Λ((𝟎,γ2d2sin2θdcosθ))dθγd.\displaystyle F_{\Gamma_{{\rm opt}}^{\rm right}}(\gamma)=\left\{\begin{array}[]{ll}0&\gamma<d\\ 1-{\rm e}^{-\frac{1}{\pi}\int_{0}^{\frac{\pi}{2}}\Lambda\left(\mathcal{B}\left(\boldsymbol{0},\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)\right){\rm d}\theta}&\gamma\geq d\end{array}.\right. (149)

for isotropic PPPs. The identity (66) holds since FΓopt(γ)=1(1FΓoptright(γ))2F_{\Gamma_{\rm opt}}\left(\gamma\right)=1-\left(1-F_{\Gamma_{\rm opt}^{\rm right}}\left(\gamma\right)\right)^{2}.

Now, we assume that Λ\Lambda is absolutely continuous with respect to the Lebesgue measure and has the unique Radon-Nikodym derivative λ:2+\lambda:\mathbb{R}^{2}\mapsto\mathbb{R}_{+}, i.e., Λ(𝒮)=𝒮λ(𝒙)𝑑𝒙\Lambda\left(\mathcal{S}\right)=\int_{\mathcal{S}}\lambda\left(\boldsymbol{x}\right)d\boldsymbol{x} for all Borel subsets 𝒮\mathcal{S} of 2\mathbb{R}^{2}. It can be seen that λ\lambda is a spherically symmetric function due to isotropy. Hence, by switching to polar coordinates, we can express FΓoptright(γ)F_{\Gamma_{{\rm opt}}^{\rm right}}(\gamma) in this case as

FΓoptright(γ)={0 if γ<d1e40π2g(γ,θ)dθ if γd,\displaystyle F_{\Gamma_{{\rm opt}}^{\rm right}}(\gamma)=\left\{\begin{array}[]{ll}0&\mbox{ if }\gamma<d\\ 1-{\rm e}^{-4\int_{0}^{\frac{\pi}{2}}g\left(\gamma,\theta\right){\rm d}\theta}&\mbox{ if }\gamma\geq d\end{array}\right., (152)

where g(γ,θ)=0γ2d2sin2θdcosθλ(ψ)ψdψg\left(\gamma,\theta\right)=\int_{0}^{\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta}\lambda\left(\psi\right)\psi{\rm d}\psi. The pdf of Γopt\Gamma_{\rm opt} is then equal to

fΓopt(γ)\displaystyle f_{\Gamma_{\rm opt}}\left(\gamma\right) =\displaystyle= ddγFΓopt(γ)\displaystyle\frac{{\rm d}}{{\rm d}\gamma}F_{\Gamma_{\rm opt}}\left(\gamma\right) (153)
=\displaystyle= 4(ddγ0π2g(γ,θ)dθ)e40π2g(γ,θ)dθ𝟣{γd}.\displaystyle 4\left(\frac{{\rm d}}{{\rm d}\gamma}\int_{0}^{\frac{\pi}{2}}g\left(\gamma,\theta\right){\rm d}\theta\right){\rm e}^{-4\int_{0}^{\frac{\pi}{2}}g\left(\gamma,\theta\right){\rm d}\theta}\mathsf{1}_{\left\{\gamma\geq d\right\}}.

g(γ,θ)g\left(\gamma,\theta\right) is a continuous function of γ\gamma and θ\theta, and its partial derivative with respect to γ\gamma

γg(γ,θ)=γ(γ2d2sin2θdcosθ)γ2d2sin2θλ(γ2d2sin2θdcosθ)\displaystyle\frac{\partial}{\partial\gamma}g\left(\gamma,\theta\right)=\frac{\gamma\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)}{\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}}\lambda\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)

is also a continuous function of γ\gamma due to continuity of λ\lambda, which can be obtained by applying Leibniz rule for differentiation under integral sign. Hence, applying Leibniz rule one more time to differentiate 0π2g(γ,θ)dθ\int_{0}^{\frac{\pi}{2}}g\left(\gamma,\theta\right){\rm d}\theta with respect to γ\gamma, we obtain

ddγ0π2g(γ,θ)dθ\displaystyle\frac{{\rm d}}{{\rm d}\gamma}\int_{0}^{\frac{\pi}{2}}g\left(\gamma,\theta\right){\rm d}\theta =\displaystyle= 0π2γg(γ,θ)dθ\displaystyle\int_{0}^{\frac{\pi}{2}}\frac{\partial}{\partial\gamma}g\left(\gamma,\theta\right){\rm d}\theta
=\displaystyle= 0π2γ(γ2d2sin2θdcosθ)γ2d2sin2θλ(γ2d2sin2θdcosθ)dθ.\displaystyle\int_{0}^{\frac{\pi}{2}}\frac{\gamma\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right)}{\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}}\lambda\left(\sqrt{\gamma^{2}-d^{2}\sin^{2}\theta}-d\cos\theta\right){\rm d}\theta.

Plugging the above expression into (153) and using the definition of the function g(γ,θ)g\left(\gamma,\theta\right), we obtain the pdf of Γopt\Gamma_{\rm opt} for isotropic PPPs as stated in Theorem 8.

Appendix I Proof of Theorem 9

The main proof idea is similar to the one given for Theorem 4. We first argue why Γopt,diff\Gamma_{\rm opt,diff} cannot be smaller than 2d(𝖲𝖭𝖱~1𝖲𝖭𝖱~2𝖲𝖭𝖱~1+𝖲𝖭𝖱~2)2d\left(\frac{\widetilde{\mathsf{SNR}}_{1}\widetilde{\mathsf{SNR}}_{2}}{\widetilde{\mathsf{SNR}}_{1}+\widetilde{\mathsf{SNR}}_{2}}\right). The optimum relay location 𝒙\boldsymbol{x}^{\star} minimizing s^diff(𝒙)\widehat{s}_{\rm diff}\left(\boldsymbol{x}\right) over 2\mathbb{R}^{2} must be located on the line segment \mathcal{L} connecting 𝒙s\boldsymbol{x}_{\rm s} and 𝒙d\boldsymbol{x}_{\rm d}. Otherwise, we can always consider the projection of 𝒙\boldsymbol{x}^{\star} on \mathcal{L}, and obtain a strictly smaller value for s^diff(𝒙)\widehat{s}_{\rm diff}\left(\boldsymbol{x}\right). At 𝒙\boldsymbol{x}^{\star}, we must also have 𝖲𝖭𝖱~1𝒙s𝒙=𝖲𝖭𝖱~2𝒙𝒙d\widetilde{{\sf SNR}}_{1}\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}^{\star}\right\|=\widetilde{{\sf SNR}}_{2}\left\|\boldsymbol{x}^{\star}-\boldsymbol{x}_{\rm d}\right\|. Otherwise, we can obtain a strictly smaller value for s^diff(𝒙)\widehat{s}_{\rm diff}\left(\boldsymbol{x}\right) in a neighborhood of 𝒙\boldsymbol{x}^{\star} by moving towards 𝒙s\boldsymbol{x}_{\rm s} or 𝒙d\boldsymbol{x}_{\rm d}. Combining these observations and using 𝒙s𝒙+𝒙𝒙d=2d\left\|\boldsymbol{x}_{\rm s}-\boldsymbol{x}^{\star}\right\|+\left\|\boldsymbol{x}^{\star}-\boldsymbol{x}_{\rm d}\right\|=2d, we obtain s^diff(𝒙)=2d𝖲𝖭𝖱~1𝖲𝖭𝖱~2𝖲𝖭𝖱~1+𝖲𝖭𝖱~2\widehat{s}_{\rm diff}\left(\boldsymbol{x}^{\star}\right)=2d\frac{\widetilde{\mathsf{SNR}}_{1}\widetilde{\mathsf{SNR}}_{2}}{\widetilde{\mathsf{SNR}}_{1}+\widetilde{\mathsf{SNR}}_{2}}.

For the rest of the proof, we will assume 𝖲𝖭𝖱~2>𝖲𝖭𝖱~1\widetilde{{\sf SNR}}_{2}>\widetilde{{\sf SNR}}_{1} without loss of generality, . The analysis for 𝖲𝖭𝖱~1>𝖲𝖭𝖱~2\widetilde{{\sf SNR}}_{1}>\widetilde{{\sf SNR}}_{2} is the same due to symmetry in the problem. The case 𝖲𝖭𝖱~2=𝖲𝖭𝖱~1\widetilde{{\sf SNR}}_{2}=\widetilde{{\sf SNR}}_{1} reduces to the equal 𝖲𝖭𝖱{\sf SNR} case analyzed in Theorem 4. We categorize the relay locations into two classes according to

Φright=Φright2andΦleft=Φleft2,\displaystyle\Phi_{\rm right}=\Phi\bigcap\mathbb{R}^{2}_{\rm right}\,\,\text{and}\,\,\Phi_{\rm left}=\Phi\bigcap\mathbb{R}^{2}_{\rm left}, (154)

where right2={(x1,x2)2:x10}\mathbb{R}^{2}_{\rm right}=\left\{\left(x_{1},x_{2}\right)^{\top}\in\mathbb{R}^{2}:x_{1}\geq 0\right\} and left2={(x1,x2)2:x1<0}\mathbb{R}^{2}_{\rm left}=\left\{\left(x_{1},x_{2}\right)^{\top}\in\mathbb{R}^{2}:x_{1}<0\right\}. Let Γopt,diffright\Gamma_{\rm opt,diff}^{\rm right} and Γopt,diffleft\Gamma_{\rm opt,diff}^{\rm left} be defined as Γopt,diffrightmin𝑿Φrights^diff(𝑿)\Gamma_{\rm opt,diff}^{\rm right}\triangleq\min_{\boldsymbol{X}\in\Phi_{\rm right}}\widehat{s}_{\rm diff}\left(\boldsymbol{X}\right) and Γopt,diffleftmin𝑿Φlefts^diff(𝑿)\Gamma_{\rm opt,diff}^{\rm left}\triangleq\min_{\boldsymbol{X}\in\Phi_{\rm left}}\widehat{s}_{\rm diff}\left(\boldsymbol{X}\right). With these definitions, we have Γopt,diff=min{Γopt,diffright,Γopt,diffleft}\Gamma_{\rm opt,diff}=\min\left\{\Gamma_{\rm opt,diff}^{\rm right},\Gamma_{\rm opt,diff}^{\rm left}\right\}. Γopt,diffright\Gamma_{\rm opt,diff}^{\rm right} and Γopt,diffleft\Gamma_{\rm opt,diff}^{\rm left} are independent but not identically distributed random variables due to scaling with different 𝖲𝖭𝖱{\sf SNR} values. Hence, the cdf of Γopt,diff\Gamma_{\rm opt,diff} is given by

FΓopt,diff(γ)=1(1FΓopt,diffright(γ))(1FΓopt,diffleft(γ)).\displaystyle F_{\Gamma_{\rm opt,diff}}\left(\gamma\right)=1-\left(1-F_{\Gamma_{\rm opt,diff}^{\rm right}}\left(\gamma\right)\right)\cdot\left(1-F_{\Gamma_{\rm opt,diff}^{\rm left}}\left(\gamma\right)\right). (155)

We observe that if 𝑿Φleft\boldsymbol{X}\in\Phi_{\rm left}, then s^diff(𝑿)=𝖲𝖭𝖱~2𝑿𝒙d\widehat{s}_{\rm diff}\left(\boldsymbol{X}\right)=\widetilde{{\sf SNR}}_{2}\left\|\boldsymbol{X}-\boldsymbol{x}_{\rm d}\right\| (i.e., 𝖲𝖭𝖱~2>𝖲𝖭𝖱~1\widetilde{{\sf SNR}}_{2}>\widetilde{{\sf SNR}}_{1}), and the analysis in the proof of Theorem 4 directly applies to obtain FΓopt,diffleft(γ)F_{\Gamma_{\rm opt,diff}^{\rm left}}\left(\gamma\right). For the cdf of Γopt,diffleft\Gamma_{\rm opt,diff}^{\rm left}, we also proceed as in the proof of Theorem 4. Consider the restriction of Φright\Phi_{\rm right} to a disc (𝟎,τ)\mathcal{B}\left(\boldsymbol{0},\tau\right) centered at the origin with radius τ\tau, which we call Φright,τ\Phi_{{\rm right},\tau}. Let NN be the number of relays in Φright,τ\Phi_{{\rm right},\tau}. Given the event {N=n}\left\{N=n\right\}, all relays will be uniformly distributed over the half disc (𝟎,τ)right2\mathcal{B}\left(\boldsymbol{0},\tau\right)\bigcap\mathbb{R}^{2}_{\rm right}. Let 𝑼\boldsymbol{U} be such a uniformly distributed relay location and define Γ=s^diff(𝑼)\Gamma=\widehat{s}_{\rm diff}\left(\boldsymbol{U}\right). Depending on the angle Θ\Theta and magnitude Ψ\Psi of 𝑼\boldsymbol{U}, we have Γ\Gamma is either equal to 𝖲𝖭𝖱~2𝑼𝒙d\widetilde{{\sf SNR}}_{2}\left\|\boldsymbol{U}-\boldsymbol{x}_{\rm d}\right\| or 𝖲𝖭𝖱~1𝑼𝒙s\widetilde{{\sf SNR}}_{1}\left\|\boldsymbol{U}-\boldsymbol{x}_{\rm s}\right\|. In particular, when Θ[θ,π2][π2,π2+θ]\Theta\in\left[\theta^{\star},\frac{\pi}{2}\right]\bigcup\left[-\frac{\pi}{2},-\frac{\pi}{2}+\theta^{\star}\right], we always have Γ=𝖲𝖭𝖱~2𝑼𝒙d\Gamma=\widetilde{{\sf SNR}}_{2}\left\|\boldsymbol{U}-\boldsymbol{x}_{\rm d}\right\|, where θ=arccos(c2+1c21)\theta^{\star}=\arccos\left(\frac{c^{2}+1}{c^{2}-1}\right) and c𝖲𝖭𝖱~2𝖲𝖭𝖱~1c\triangleq\frac{\widetilde{{\sf SNR}}_{2}}{\widetilde{{\sf SNR}}_{1}}. For Θ(θ,θ)\Theta\in\left(-\theta^{\star},\theta^{\star}\right), we have Γ=𝖲𝖭𝖱~2𝑼𝒙d\Gamma=\widetilde{{\sf SNR}}_{2}\left\|\boldsymbol{U}-\boldsymbol{x}_{\rm d}\right\| when Ψ[0,r1(Θ)][r2(Θ),τ]\Psi\in\left[0,r_{1}\left(\Theta\right)\right]\bigcup\left[r_{2}\left(\Theta\right),\tau\right] and Γ=𝖲𝖭𝖱~1𝑼𝒙s\Gamma=\widetilde{{\sf SNR}}_{1}\left\|\boldsymbol{U}-\boldsymbol{x}_{\rm s}\right\| when Ψ(r1(Θ),r2(Θ))\Psi\in\left(r_{1}\left(\Theta\right),r_{2}\left(\Theta\right)\right), where the functions r1(θ)r_{1}\left(\theta\right) and r2(θ)r_{2}\left(\theta\right) are defined as r1(θ)=c2+1c21dcos(θ)d(c2+1c21)2cos2(θ)1r_{1}\left(\theta\right)=\frac{c^{2}+1}{c^{2}-1}d\cos\left(\theta\right)-d\sqrt{\left(\frac{c^{2}+1}{c^{2}-1}\right)^{2}\cos^{2}\left(\theta\right)-1} and r2(θ)=c2+1c21dcos(θ)+d(c2+1c21)2cos2(θ)1r_{2}\left(\theta\right)=\frac{c^{2}+1}{c^{2}-1}d\cos\left(\theta\right)+d\sqrt{\left(\frac{c^{2}+1}{c^{2}-1}\right)^{2}\cos^{2}\left(\theta\right)-1}.666Here, we assume that τc2+1c21d+d(c2+1c212)1\tau\geq\frac{c^{2}+1}{c^{2}-1}d+d\sqrt{\left(\frac{c^{2}+1}{c^{2}-1}^{2}\right)-1}. Analyzing these different cases, averaging over NN, taking the limit τ\tau\rightarrow\infty and using (155), we arrive at (91).

References

  • [1] S. Atapattu, H. Inaltekin, and J. Evans, “Location-based optimum relay selection in random spatial networks,” in Proc. IEEE Int. Conf. Commun. (ICC), Shanghai, China, May 2019.
  • [2] H. Inaltekin, S. Atapattu, and J. Evans, “Limited-feedback distributed relay selection for random spatial wireless networks,” in Proc. IEEE Global Telecommn. Conf. (GLOBECOM), Hawaii, United States, Dec. 2019.
  • [3] D. Tse and P. Viswanath, Fundamentals of Wireless Communication.   New York, NY, USA: Cambridge University Press, 2005.
  • [4] R. Pabst, B. H. Walke, D. C. Schultz, P. Herhold, H. Yanikomergolu, S. Mukherjee, H. Viswanathan, M. Lott, W. Zirwas, M. Dohler, H. Aghvami, D. D. Falconer, and G. P. Fettweis, “Relay-based deployment concepts for wireless and mobile broadband radio,” IEEE Commun. Mag., vol. 42, no. 9, pp. 80–89, Sep. 2004.
  • [5] A. Nosratinia, T. E. Hunter, and A. Hedayat, “Cooperative communication in wireless networks,” IEEE Commun. Mag., vol. 42, no. 10, pp. 74–80, Oct 2004.
  • [6] O. Somekh, O. Simeone, H. V. Poor, and S. Shamai, “Cellular systems with non-regenerative relaying and cooperative base stations,” IEEE Trans. Wireless Commun., vol. 9, no. 8, pp. 2654–2663, Aug. 2010.
  • [7] L. Li, H. V. Poor, and L. Hanzo, “Non-coherent successive relaying and cooperation: Principles, designs and applications,” IEEE Commun. Surveys Tuts., vol. 17, no. 3, pp. 1708–1737, Third Quarter 2015.
  • [8] E. C. van der Meulen, “Three-terminal communication channels,” Adv. Appl. Probab., vol. 3, no. 1, pp. 120–154, 1971.
  • [9] T. M. Cover and A. A. E. Gamal, “Capacity theorems for the relay channel,” IEEE Trans. Inform. Theory, vol. IT-25, no. 5, pp. 572–584, Sep. 1979.
  • [10] B. Schein and R. Gallager, “The gaussian parallel relay network,” in Proc. IEEE Int. Symp. Information Theory (ISIT), Sorrento, Italy, Jun. 2000, p. 22.
  • [11] J. N. Laneman and G. W. Wornell, “Distributed space-time coded protocols for exploiting cooperative diversity in wireless networks,” IEEE Trans. Inform. Theory, vol. 49, no. 10, pp. 2415–2425, Oct. 2003.
  • [12] P. Gupta and P. R. Kumar, “Towards an information theory of large networks: an achievable rate region,” IEEE Trans. Inform. Theory, vol. 49, no. 8, pp. 1877–1894, Aug. 2003.
  • [13] A. Sendonaris, E. Erkip, and B. Aazhang, “User cooperation diversity - Part I: System description,” IEEE Trans. Commun., vol. 51, no. 11, pp. 1927–1938, Nov 2003.
  • [14] ——, “User cooperation diversity - Part II: Implementation aspects and performance analysis,” IEEE Trans. Commun., vol. 51, no. 11, pp. 1939–1948, Nov 2003.
  • [15] A. Reznik, S. R. Kulkarni, and S. Verdu, “Degraded Gaussian multirelay channel: Capacity and optimal power allocation,” IEEE Trans. Inform. Theory, vol. 50, no. 12, pp. 3037–3046, Dec. 2004.
  • [16] J. N. Laneman, D. N. C. Tse, and G. W. Wornell, “Cooperative diversity in wireless networks: Efficient protocols and outage behavior,” IEEE Trans. Inform. Theory, vol. 50, no. 12, pp. 3062–3080, Dec. 2004.
  • [17] L.-L. Xie and P. R. Kumar, “A network information theory for wireless communication: Scaling laws and optimal operation,” IEEE Trans. Inform. Theory, vol. 50, no. 5, pp. 748–767, May 2004.
  • [18] ——, “An achievable rate for the multiple-level relay channel,” IEEE Trans. Inform. Theory, vol. 51, no. 4, pp. 1348–1358, Apr. 2005.
  • [19] G. Kramer, M. Gastpar, and P. Gupta, “Cooperative strategies and capacity theorems for relay networks,” IEEE Trans. Inform. Theory, vol. 51, no. 9, pp. 3037–3063, Sep. 2005.
  • [20] A. Høst-Madsen and J. Zhang, “Capacity bounds and power allocation for wireless relay channels,” IEEE Trans. Inform. Theory, vol. 51, no. 9, pp. 3037–3063, Sep. 2005.
  • [21] B. Wang, J. Zhang, and A. Høst-Madsen, “On the capacity of MIMO relay channels,” IEEE Trans. Inform. Theory, vol. 51, no. 1, pp. 29–43, Jan. 2005.
  • [22] A. Høst-Madsen, “Capacity bounds for cooperative diversity,” IEEE Trans. Inform. Theory, vol. 52, no. 5, pp. 1522–1544, Apr. 2006.
  • [23] M. Katz and S. Shamai, “Cooperative schemes for a source and an occasional nearby relay in wireless networks,” IEEE Trans. Inform. Theory, vol. 55, no. 11, pp. 5138–5160, Nov. 2009.
  • [24] A. S. Avestimehr and D. N. C. Tse, “Outage capacity of the fading relay channel in the low-SNR regime,” IEEE Trans. Inform. Theory, vol. 53, no. 4, pp. 1401–1415, Apr. 2007.
  • [25] A. Jain, S. R. Kulkarni, and S. Verdu, “Energy efficiency of decode-and-forward for wideband wireless multicasting,” IEEE Trans. Inform. Theory, vol. 57, no. 12, pp. 7695–7713, Dec. 2011.
  • [26] R. Kolte, U. Niesen, and P. Gupta, “Energy efficient communication over the unsynchronized gaussian diamond network,” IEEE Trans. Inform. Theory, vol. 60, no. 12, pp. 7719–7731, Dec. 2014.
  • [27] A. S. Avestimehr, S. N. Diggavi, and D. N. C. Tse, “Wireless network information flow: A deterministic approach,” IEEE Trans. Inform. Theory, vol. 57, no. 4, pp. 1872–1905, Apr. 2011.
  • [28] R. Kolte, A. Ozgur, and A. E. Gamal, “Capacity approximations for gaussion relay networks,” IEEE Trans. Inform. Theory, vol. 61, no. 9, pp. 4721–4734, Sep. 2015.
  • [29] P. Rost, G. Fettweis, and J. N. Laneman, “Energy- and cost-efficient mobile communication using multi-cell MIMO and relaying,” IEEE Trans. Wireless Commun., vol. 11, no. 9, pp. 3377–3387, Sep. 2012.
  • [30] Third Generation Partnership Project (3GPP), “Evolved universal terrestrial radio access (E-UTRA): Physical layer for relaying operation,” Technical Report 3GPP TS 36.216 V15.0.0, Jun. 2018.
  • [31] ——, “Evolved universal terrestrial radio access (E-UTRA): Further advancements for E-UTRA physical layer aspects,” Technical Report 3GPP TR 36.184 V9.2.0, Mar. 2017.
  • [32] M. Thakur, N. Fawaz, and M. Medard, “Optimal relay location and power allocation for low SNR broadcast relay channels,” in Proc. IEEE INFOCOM, Apr. 2011, pp. 2822–2830.
  • [33] W. Guo and T. O’Farrell, “Relay deployment in cellular networks: Planning and optimization,” IEEE J. Select. Areas Commun., vol. 31, no. 8, pp. 1597–1606, Aug. 2013.
  • [34] C. Nazaroglu, A. Ozgur, and C. Fragouli, “Wireless network simplification: The Gaussian nn-relay diamond network,” IEEE Trans. Inform. Theory, vol. 60, no. 10, pp. 6329–6341, Oct 2014.
  • [35] A. Chattopadhyay, A. Sinha, M. Coupechoux, and A. Kumar, “Deploy-as-you-go wireless relay placement: An optimal sequential decision approach using the multi-relay channel model,” IEEE Trans. Mobile Comput., vol. 16, no. 2, pp. 341–354, Feb. 2017.
  • [36] G. Liu, F. R. Yu, H. Ji, and V. C. M. Leung, “In-band full-duplex relaying: A survey, research issues and challenges,” IEEE Commun. Surveys Tuts., vol. 17, no. 2, pp. 500–524, Second Quarter 2015.
  • [37] R. G. Gallager, Principles and Digital Communication.   New York, NY, USA: Cambridge University Press, 2008.
  • [38] M. Haenggi, “The meta distribution of the SIR in Poisson bipolar and cellular networks,” IEEE Trans. Wireless Commun., vol. 15, no. 4, pp. 2577–2589, Apr. 2016.
  • [39] ——, “Efficient calculation of meta distributions and the performance of user percentiles,” IEEE Wireless Commun. Lett., vol. 7, no. 6, pp. 982–985, Dec. 2018.
  • [40] ——, “SIR meta distribution of kk-tier downlink heterogeneous cellular networks with cell range expansion,” IEEE Trans. Commun., vol. 67, no. 4, pp. 3069–3081, Apr. 2019.
  • [41] R. Durrett, Probability: Theory and Examples, 2nd ed.   Belmont, CA, USA: Duxbury Press, 1996.
  • [42] S. Cho, W. Choi, and K. Huang, “QoS provisioning relay selection in random relay networks,” IEEE Trans. Veh. Technol., vol. 60, no. 6, pp. 2680–2689, Jul. 2011.
  • [43] M. Mohammadi and H. A. S. and, “Outage probability of wireless ad hoc networks with cooperative relaying,” in Proc. IEEE Global Telecommn. Conf. (GLOBECOM), Dec. 2012, pp. 4410–4416.
  • [44] A. Behnad, A. M. Rabiei, and N. C. Beaulieu, “Performance analysis of opportunistic relaying in a poisson field of amplify-and-forward relays,” IEEE Trans. Commun., vol. 61, no. 1, pp. 97–107, Jan. 2013.
  • [45] A. Altieri, L. R. Vega, P. Piantanida, and C. G. Galarza, “Analysis of a cooperative strategy for a large decentralized wireless network,” IEEE/ACM Trans. Netw., vol. 22, no. 4, pp. 1039–1051, Aug. 2014.
  • [46] A. Tukmanov, S. Boussakta, Z. Ding, and A. Jamalipour, “Outage performance analysis of imperfect-CSI-based selection cooperation in random networks,” IEEE Trans. Commun., vol. 62, no. 8, pp. 2747–2757, Aug. 2014.
  • [47] Y. Zhou and W. Zhuang, “Performance analysis of cooperative communication in decentralized wireless networks with unsaturated traffic,” IEEE Trans. Wireless Commun., vol. 15, no. 5, pp. 3518–3530, May 2016.
  • [48] H. Elkotby and M. Vu, “Outage performance of uplink user-assisted relaying in 5G cellular networks,” in Proc. IEEE Global Telecommn. Conf. (GLOBECOM), Dec. 2015.
  • [49] H. Elkotby and M. Vu, “Uplink user-assisted relaying in cellular networks,” IEEE Trans. Wireless Commun., vol. 14, no. 10, pp. 5468–5483, Oct. 2015.
  • [50] I. Krikidis, “Simultaneous information and energy transfer in large-scale networks with/without relaying,” IEEE Trans. Commun., vol. 62, no. 3, pp. 900–912, Mar. 2014.
  • [51] K. Belbase, Z. Zhang, H. Jiang, and C. Tellambura, “Coverage analysis of millimeter wave decode-and-forward networks with best relay selection,” IEEE Access, vol. 6, pp. 22 670–22 683, 2018.
  • [52] Y. Jing and H. Jafarkhani, “Single and multiple relay selection schemes and their achievable diversity orders,” IEEE Trans. Wireless Commun., vol. 8, no. 3, pp. 1414–1423, Mar. 2009.
  • [53] S. Atapattu, Y. Jing, H. Jiang, and C. Tellambura, “Relay selection and performance analysis in multiple-user networks,” IEEE J. Select. Areas Commun., vol. 31, pp. 1517–1529, Aug. 2013.
  • [54] S. H. Lim, Y. Kim, A. El Gamal, and S. Chung, “Noisy network coding,” IEEE Trans. Inform. Theory, vol. 57, no. 5, pp. 3132–3152, May 2011.
  • [55] A. S. Avestimehr, S. N. Diggavi, and D. N. C. Tse, “Wireless network information flow: A deterministic approach,” IEEE Trans. Inform. Theory, vol. 57, no. 4, pp. 1872–1905, Apr. 2011.
  • [56] M. Cardone, D. Tuninetti, R. Knopp, and U. Salim, “Gaussian half-duplex relay networks: Improved constant gap and connections with the assignment problem,” IEEE Trans. Inform. Theory, vol. 60, no. 6, pp. 3559–3575, Jun. 2014.
  • [57] Y. H. Ezzeldin, M. Cardone, C. Fragouli, and D. Tuninetti, “Network simplification in half-duplex: Building on submodularity,” IEEE Trans. Inform. Theory, vol. 65, no. 10, pp. 6801–6818, Oct. 2019.
  • [58] M. E. Eltayeb, K. Elkhalil, H. R. Bahrami, and T. Y. Al-Naffouri, “Opportunistic relay selection with limited feedback,” IEEE Trans. Commun., vol. 63, no. 8, pp. 2885–2898, Aug 2015.
  • [59] K. Elkhalil, M. E. Eltayeb, A. Kammoun, T. Y. Al-Naffouri, and H. R. Bahrami, “On the feedback reduction of multiuser relay networks using compressive sensing,” IEEE Trans. Commun., vol. 64, no. 4, pp. 1437–1450, Apr. 2016.
  • [60] H. Wang, “Full-diversity uncoordinated cooperative transmission for asynchronous relay networks,” IEEE Trans. Veh. Technol., vol. 66, no. 1, pp. 468–480, Jan. 2017.
  • [61] M. R. Javan, N. Mokari, F. Alavi, and A. Rahmati, “Resource allocation in decode-and-forward cooperative communication networks with limited rate feedback channel,” IEEE Trans. Veh. Technol., vol. 66, no. 1, pp. 256–267, Jan. 2017.
  • [62] M. Haenggi, “On routing in random rayleigh fading networks,” IEEE Trans. Wireless Commun., vol. 4, no. 4, pp. 1553–1562, Jul. 2005.
  • [63] M. Haenggi, “On distances in uniformly random networks,” IEEE Transactions on Information Theory, vol. 51, no. 10, pp. 3584–3586, 2005.
  • [64] M. Sikora, J. N. Laneman, M. Haenggi, D. J. Costello, and T. E. Fuja, “Bandwidth- and power-efficient routing in linear wireless networks,” IEEE Transactions on Information Theory, vol. 52, no. 6, pp. 2624–2633, 2006.
  • [65] J. G. Andrews, S. Weber, M. Kountouris, and M. Haenggi, “Random access transport capacity,” IEEE Trans. Wireless Commun., vol. 9, no. 6, pp. 2101–2111, Jun. 2010.
  • [66] R. Vaze, “Throughput-delay-reliability tradeoff with ARQ in wireless Ad Hoc networks,” IEEE Transactions on Wireless Communications, vol. 10, no. 7, pp. 2142–2149, 2011.
  • [67] Y. Chen and J. G. Andrews, “An upper bound on multihop transmission capacity with dynamic routing selection,” IEEE Transactions on Information Theory, vol. 58, no. 6, pp. 3751–3765, 2012.
  • [68] K. Stamatiou and M. Haenggi, “Delay characterization of multihop transmission in a Poisson field of interference,” IEEE/ACM Transactions on Networking, vol. 22, no. 6, pp. 1794–1807, 2014.
  • [69] B. Błaszczyszyn and P. Mühlethaler, “Random linear multihop relaying in a general field of interferers using spatial Aloha,” IEEE Transactions on Wireless Communications, vol. 14, no. 7, pp. 3700–3714, 2015.
  • [70] Y. Liang and T. Li, “End-to-end throughput in multihop wireless networks with random relay deployment,” IEEE Transactions on Signal and Information Processing over Networks, vol. 4, no. 3, pp. 613–625, 2018.
  • [71] T. Yuan, LTE-Advanced Relay Technology and Standardization.   Berlin, Germany: Springer-Verlag, 2013.
  • [72] M. Iwamura, H. Takahashi, and S. Nagata, “Relay technology in LTE-Advanced,” NTT DOCOMO Technical Journal, vol. 12, no. 2, pp. 29–36, 2010.
  • [73] A. Goldsmith, Wireless Communications.   New York, NY, USA: Cambridge University Press, 2005.
  • [74] L. H. Ozarow, S. Shamai, and A. D. Wyner, “Information theoretic consideration for cellular mobile radio,” IEEE Trans. Veh. Technol., vol. 43, no. 2, pp. 359–378, May. 1994.
  • [75] E. Ozfatura, O. Ercetin, and H. Inaltekin, “Optimal network-assisted multi-user DASH video streaming,” IEEE Trans. Broadcast., vol. 64, no. 2, pp. 247–265, Jun. 2018.
  • [76] J. F. C. Kingman, Poisson Processes.   Oxford, UK: Clarendon Press, 1993.
  • [77] P. Billingsley, Probability and Measure, 3rd ed.   New York, NY, USA: John Wiley & Sons, 1995.
  • [78] S. Ak, H. Inaltekin, and H. V. Poor, “Gaussian approximation for the downlink interference in heterogenous cellular networks,” in Proc. 2016 IEEE International Symposium on Information Theory (ISIT), Barcelona, Spain, Jul. 2016, pp. 1611–1615.
  • [79] H. Inaltekin, M. Chiang, H. V. Poor, and S. Wicker, “On unbounded path-loss models: Effects of singularity on wireless network performance,” IEEE J. Sel. Areas Commun., vol. 27, no. 7, pp. 1078–1092, Sep. 2009.
  • [80] S. Ak, H. Inaltekin, and H. V. Poor, “Downlink outage performance of heterogenous cellular networks,” in Proc. 2016 IEEE International Symposium on Information Theory (ISIT), Barcelona, Spain, Jul. 2016, pp. 1616–1620.
  • [81] H. Inaltekin, “Gaussian approximation for the wireless multi-access interference distribution,” IEEE Trans. Signal Process., vol. 60, no. 11, pp. 6114–6120, Nov. 2012.
  • [82] T. Samarasinghe, H. Inaltekin, and J. S. Evans, “On the outage capacity of opportunistic beamforming with random user locations,” IEEE Trans. Commun., vol. 62, no. 8, pp. 3015–3026, Aug. 2014.
  • [83] S. Ak, H. Inaltekin, and H. V. Poor, “A tractable framework for the analysis of dense heterogenous cellular networks,” IEEE Trans. Commun., vol. 66, no. 7, pp. 3151–3171, Jul. 2018.
  • [84] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions, With Formulas, Graphs, and Mathematical Tables,.   New York, NY, USA: Dover Publications, Inc., 1974.
  • [85] T. Samarasinghe, H. Inaltekin, and J. S. Evans, “The feedback-capacity tradeoff for opportunistic beamforming,” in Proc. 2011 IEEE International Conference on Communications (ICC), Kyoto, Japan, Jun. 2011, pp. 1–6.
  • [86] H. Inaltekin, T. Samarasinghe, and J. S. Evans, “Rate optimal limited feedback policies for the mimo downlink,” in Proc. 2011 International Symposium on Modeling and Optimization in Mobile, Ad Hoc, and Wireless Networks, Princeton, NJ, US, May 2011, pp. 294–299.
  • [87] T. Samarasinghe, H. Inaltekin, and J. S. Evans, “Vector broadcast channels: Optimality of threshold feedback policies,” in Proc. 2011 IEEE International Symposium on Information Theory (ISIT), Saint Petersburg, Russia, Jul.-Aug. 2011, pp. 1292–1296.
  • [88] T. Samarasinghe, H. Inaltekin, and J. S.Evans, “Vector broadcast channels: Optimal threshold selection problem,” in Proc. 2011 IEEE International Symposium on Information Theory (ISIT), Saint Petersburg, Russia, Jul.-Aug. 2011, pp. 1906–1910.
  • [89] T. Samarasinghe, H. Inaltekin, and J. S. Evans, “Optimal selective feedback policies for opportunistic beamforming under peak feedback constraints,” in Proc. 2012 IEEE International Symposium on Information Theory (ISIT), Boston, MA, US, Jul. 2012, pp. 2919–2923.
  • [90] T. Samarasinghe, H. Inaltekin, and J. S.Evans, “Optimal selective feedback policies for opportunistic beamforming,” IEEE Trans. Inf. Theory, vol. 59, no. 5, pp. 2897–2913, May 2013.
  • [91] T. Samarasinghe, H. Inaltekin, and J. S. Evans, “Optimal user selection and the feedback-capacity tradeoff for opportunistic beamforming,” ELSEVIER Performance Evaluation, vol. 70, no. 7-8, pp. 472–492, Jul. 2013.
  • [92] A. AlAmmouri, J. G. Andrews, and F. Baccelli, “SINR and throughput of dense cellular networks with stretched exponential path loss,” IEEE Trans. Wireless Commun., vol. 17, no. 2, pp. 1147–1160, Feb. 2018.
  • [93] M. Haenggi, Stochastic Geometry for Wireless Networks.   New York, NY: Cambridge University Press, 2013.
  • [94] E. Koyuncu, M. Shabanighazikelayeh, and H. Seferoglu, “Deployment and trajectory optimization of UAVs: A quantization theory approach,” IEEE Trans. Wireless Commun., vol. 17, no. 12, pp. 8531–8546, Dec. 2018.
  • [95] M. Shabanighazikelayeh and E. Koyuncu, “Outage-optimized deployment of UAVs,” in Proc. IEEE 30th Annual International Symposium on Personal, Indoor and Mobile Radio Communications (PIMRC), Istanbul, Turkey, 2019, pp. 1–6.
  • [96] E. W. Ng and M. Geller, “A table of integrals of the error functions,” J. Res. Nat. Bureau Standards-B, vol. 73B, pp. 1–20, Jan.–Mar. 1969.
  • [97] K.-T. Fang, S. Kotz, and K. W. Ng, Symmetric Multivariate and Related Distributions.   New York, NY, USA: Springer-Science+Business Media, 1990.